Advances in Colloid and Interface Science

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

CIS-01858; No of Pages 20

Advances in Colloid and Interface Science xxx (2018) xxx–xxx

Contents lists available at ScienceDirect

Advances in Colloid and Interface Science

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

Historical perspective

The application of atomic force microscopy in mineral flotation


Yaowen Xing a,b,c, Mengdi Xu a, Xiahui Gui b,c, Yijun Cao b,d, Bent Babel e, Martin Rudolph e, Stefan Weber c,
Michael Kappl c, Hans-Jürgen Butt c,⁎
a
School of Chemical Engineering and Technology, China University of Mining and Technology, Xuzhou 221116, China
b
Chinese National Engineering Research Center of Coal Preparation and Purification, China University of Mining and Technology, Xuzhou 221116, China
c
Max Planck Institute for Polymer Research, Ackermannweg 10, 55128 Mainz, Germany
d
Henan Province Industrial Technology Research Institute of Resources and Materials, Zhengzhou University, Zhengzhou 450001, China
e
Helmholtz-Zentrum Dresden-Rossendorf, Helmholtz Institute Freiberg for Resource Technology, Chemnitzer Str. 40, 09599 Freiberg, Germany

a r t i c l e i n f o a b s t r a c t

Available online xxxx During the past years, atomic force microscopy (AFM) has matured to an indispensable tool to characterize
nanomaterials in colloid and interface science. For imaging, a sharp probe mounted near to the end of a cantilever
Keywords: scans over the sample surface providing a high resolution three-dimensional topographic image. In addition, the
Atomic force microscopy AFM tip can be used as a force sensor to detect local properties like adhesion, stiffness, charge etc. After the inven-
Mineral flotation tion of the colloidal probe technique it has also become a major method to measure surface forces. In this review,
Surface imaging
we highlight the advances in the application of AFM in the field of mineral flotation, such as mineral morphology
Inter-particle force
Bubble-particle interaction
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 and
distance involving deformable bubble as a force sensor will further assist the fundamental understanding of flo-
tation mechanism.
© 2018 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2. Imaging of minerals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3. Water at mineral surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.1. Water structure near the interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.2. Surface nanobubbles characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4. Reagent adsorption on mineral surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.1. Structure of adsorbed reagent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.2. Single molecule force spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5. Quantification of inter-particle force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6. Bubble-particle interaction and thin liquid film drainage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
7. Other applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
8. Conclusions and perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0

1. Introduction

Froth flotation, belonging to the family of heterocoagulation separa-


⁎ Corresponding author.
tion techniques, is based on the difference in surface hydrophobicity of
E-mail addresses: guixiahui1985@163.com (X. Gui), yijuncao@126.com (Y. Cao), dispersed particles. It has been widely used in mineral processing [1–4],
kappl@mpip-mainz.mpg.de (M. Kappl), butt@mpip-mainz.mpg.de (H.-J. Butt). fine coal upgrading [5–7], wastewater treatment [8,9], oil sands

https://doi.org/10.1016/j.cis.2018.01.004
0001-8686/© 2018 Elsevier B.V. All rights reserved.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
2 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

processing [10,11], fly ash decarburization [12–14] and pulp-deinking surface usually an instability accors once a certain distance has been
[15,16]. For mineral flotation, when the raw minerals are collected reached [41,42].
from the underground, crushing and grinding is required for liberating With the invention of Atomic force microscopy (AFM) a new tool be-
the valuable components from the interlocking particles. Then, these came available for measuring the interaction between particles and
fine particles are mixed with water and conditioned with appropriate bubbles and for imaging the surface structure of minerals. (Fig. 1).
reagents (collectors, frothers, depressants and regulators). Air is finally AFM was first developed by Binnig et al. [43] in 1986, for imaging the
introduced and hydrophobic particles are captured by the rising bubbles surface topography at nanometer resolution [44]. A sharp probe
while hydrophilic particles remain in the pulp. mounted near the end of a cantilever is raster scanned along the sample
The essential sub-processes in flotation are: wetting, reagents ad- surface. A laser beam was used to monitor the deflection of the cantile-
sorption on minerals, inter-particle force, and bubble-particle interac- ver via a position sensitive diode detector (PSD). Cantilever deflection is
tion. Each step plays a critical role in the final flotation recovery. A proportional to the force acting on the tip. Usually a feedback loop is
comprehensive understanding of these sub-processes is of great funda- used to keep the force between tip and sample constant by moving
mental and practical importance to design high-efficiency flotation the sample up and down during scanning. In that case the height of
process. Up to now, a number of studies have been carried out to inves- the sample is plotted versus the lateral position to obtain a three-
tigate these sub-processes [17–28]. Contact angle measurement is the dimensional topographic image both in air or liquid condition [45–47].
most commonly used method in the study of mineral-water wetting Typically, three common imaging modes, i.e., contact mode, tapping
[29]. Spectroscopic techniques such as infrared spectroscopy, X-ray mode, and non-contact mode, are available in commercial AFMs based
photoelectron spectroscopy and sum-frequency generation spectros- on the way in which the tip scans over the sample [47,48]. In contact
copy (SFG) are usually applied to better understand reagent adsorption mode, the tip remains in contact with the sample surface by applying
on mineral surface and water structure at mineral-water interface a constant load between tip and the sample (Fig. 1. b). Due to the
[17–21]. Zeta potentials are measured to characterize the charge on par- large contact pressure, the tip or the sample may get damaged during
ticles which is essential to predict inter-particle electrostatic interaction. the imaging, leading to the decrease in resolution [48,49]. For the soft
Theory and computational fluid dynamics are reported to study hydro- surface, contact mode imaging will underestimate the height of the
dynamic inter-particle or bubble-particle interaction [22–28]. However, sample due to the deformation effect. For tapping mode, the tip is oscil-
the flotation mechanisms at the nanoscale are still not well understood lated with a value close to its resonance frequency. The tip comes into
due to the difficulty in experimental verification. and out of contact with the sample repeatedly at the lower turning
Particularly, surface and interfacial forces (van der Waals, electrical point. This intermittent contact between tip and sample leads to a
double layer, hydrophobic force, hydration, hydrodynamic, and adhe- damping of the tip oscillation, reflected both by a reduction in ampli-
sion forces) are of great importance to understand inter-particle or tude and a shift of the resonance frequency. Depending on the feedback
bubble-particle interaction in flotation cells [1,30]. According to the used during intermittent mode, one discriminates between amplitude
classical Derjaguin-Landau-Verwey-Overbeek (DLVO) theory, the sta- modulation (AM) and frequency modulation (FM). In the more com-
bility of a colloidal system is controlled by two components [30,31], mon AM mode, the tip is approached to the sample until a preset ampli-
namely van der Waals dispersion and electrostatic double layer forces. tude reduction of cantilever oscillation is reached, and this reduced
The van der Waals force between different bodies depends on their amplitude is kept constant during scanning using a feedback loop. In
differences in optical properties and can be characterized by the FM mode, a preselected shift of the resonance frequency is used for
Hamaker constant. For the interaction of a solid/liquid with a liquid/ the feedback during approach and scanning. FM mode was originally
gas interface, the van der Waals force is repulsive; it favors thickening developed for AFM in ultra-high vacuum (UHV), where AM mode suf-
of the liquid film. For the electrostatic double layer force, surface charges fers from a slow response time of the feedback due to a much higher
and the salt concentration are the key parameters for determining its quality factor of the oscillating cantilever, as damping by a fluid is miss-
value. ing [50]. Recently it was demonstrated, that FM mode can also be used
In addition, non-DLVO forces such as hydration and hydrophobic in liquids [51], allowing imaging and force measurements with atomic
forces are present in flotation system. A short-ranged repulsive hydra- resolution. Compared with contact mode, intermittent contact mode
tion force emerges when two hydrophilic surfaces are brought into con- allows to image soft surface with a higher resolution, as interaction
tact and the structure of water is required to change upon approach between tip and sample is minimized and especially lateral shear is
[32–34]. In contrast, between two hydrophobic surfaces attractive eliminated.
forces are observed [35–38]. Hydrophobic forces are usually the driving Soon after its introduction, AFM was used as a force sensor to mea-
force for the attachment of solid particles to bubbles. Laskowski and sure forces, including surface forces, adhesion force, and hydrodynamic
Kitchener [39] first found that the water films present on methylated force, between different surfaces or molecules of interest, especially
silica surfaces were unstable and the film ruptured spontaneously. with the aid of the so called “colloidal probe technique” [47,52–54].
Their experiment indicates that in addition to the traditional DLVO The force is obtained by multiplying the measured probe deflection by
forces, an attractive force also acts between the bubbles and the hydro- the spring constant of the cantilever. In the colloid probe technique, a
phobic solid particles. Three years later, Blake and Kitchener [40] exper- micrometer-size particle is attached to the end of the cantilever and re-
imentally measured the thickness of the water film when a small gas places the sharp tip. Usually the particle is spherical to facilitate the
bubble was slowly advanced towards a polished silica plate with differ- quantitative analysis. Thus, the interaction geometry between the
ent hydrophobicity. They showed that the film rupture thickness on probe and surface is greatly simplified.
methylated silica ranged from 60 to 220 nm. In this case, the thin AFM has also been widely used by flotation scientists to imaging
water film on the methylated silica surface was unstable due to a mineral morphology and measure the force between single particles.
long-range attractive force. The first direct evidence for long-range at- However, it is difficult to predict flotation recovery directly from force
tractive hydrophobic forces between solid surfaces was provided by measurements because a large number of particles are involved in
Pashley and Israelachvili using the Surface Force Apparatus (SFA) flotation practice. Still, force measurements can provide useful input
[35,36]. In SFA, the force between two hydrophobized mica surfaces parameters for computational fluid dynamics simulation to design
was measured. To date, attractive hydrophobic forces have been ac- high efficiency flotation process.
cepted as the reason for bubble-particle attachment, although its origin The aim of this review is to demonstrate the versatility, flexibility,
remains under debate. Measuring the hydrophobic force between a and future potential of AFM in order to shed new light on the flotation
bubble and a particle versus distance is, however, challenging. Due to mechanisms. We highlight the current state of the art in the application
its strong attractive characteristics and the softness of the bubble of AFM in the field of mineral flotation, such as mineral morphology

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx 3

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.

imaging, water at mineral surface, reagent adsorption, inter-particle faces of phyllosilicate (kaolinite). Negatively charged glass or mica and
force, and bubble-particle interaction. Limitations and future positively charged alumina were used as the substrates for kaolinite de-
perspectives are also discussed. This review is presented in a way that position. Due to the electrostatic double-layer attraction, the positively
is accessible and valuable to flotation researchers with basic knowledge charged alumina face of kaolinite particles attached to the glass or
background regarding AFM. mica surface, thus exposing the silica faces. In contrast, the alumina
face of kaolinite particles was exposed, when alumina was used as a
2. Imaging of minerals substrate. It was found that tetrahedral oxygen atoms on the silica
face formed a closed hexagonal ring-like network with a vacancy in
AFM was used to image pure mineral surfaces such as mica, galena, the centre. On the alumina face, the hexagonal lattice ring of hydroxyls
molybdenite and phyllosilicate. All can be easily cleaved [55–61], to surrounded a hydroxyl. The atomic spacing between neighboring
prepare clean, atomically flat surfaces. For example, Gupta et al. hydroxyls was determined as 0.36 ± 0.04 nm. Siretanu et al. [61]
[59,62] used AFM for imaging the crystal lattice of silica and alumina used AM-AFM to image the basal (001) planes of various types of

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
room temperature equilibrated in ultrapure water. Scale bars, 2 nm. Top insets represent zooms of atomic scale images after processing with Fourier transformation superimposed
with X-ray crystal structure of clays along the ab plane.
Reprinted with permission from [61], Copyright 2016, The Royal Society of Chemistry.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
4 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

phyllosilicates, namely gibbsite, kaolinite, illite, and Na-montmorillonite, AFM also has the ability to characterize the pore structure on coal
in water (Fig. 2). All images show the characteristic hexagonal structure surfaces [67]. Dun et al. [68] found that magmatic intrusions have a
with a periodicity ≈0.5 nm that arises from the quasi-hexagonal arrange- great impact on the pore structures of coal. The pore size of low-rank bi-
ment of silica-tetrahedra and alumina-octahedra, the classical building tuminous coal was much larger than high-rank anthracite. As a result,
units of all clay materials. the small molecular weight alkanes (collectors) are easily lost in the
In sulfide mineral flotation the oxidation state is of great importance. pores due to the capillary effect [6]. Indeed, a higher concentration of
The floatability of sulfide minerals is directly related to the degree of its collector is usually required for low-rank coal flotation than that of
surface oxidation [63]. For example, fresh galena is hydrophilic. How- high-rank anthracite.
ever, it was found that slight surface oxidation increased the hydropho- Both the sharpness and the aspect ratio of the probe influence im-
bicity and floatability of galena due to the formation of sulphur rich ages obtained with an AFM due to the convolution and broadening ef-
layers [60,64]. Therefore, collectorless flotation of sulfide minerals fect. Only probes with a high aspect ratio and small probe radius are
becomes possible by adjusting the surface oxidation. Xie et al. [57] suitable to quantitatively characterize the roughness and pore structure
studied the effect of applied electrochemical potential, with regard to on mineral surfaces.
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 3. Water at mineral surface
significantly when the potential exceeded +300 mV (Fig. 3). This was
due to the agglomeration of electrochemical oxidation products on 3.1. Water structure near the interface
the galena surface. Hampton et al. [60] studied the effect of applied
potential on electrochemical oxidation of the galena surface at pH 4.5 AFM is capable of imaging mineral surfaces in liquid. In flotation,
(Fig. 4). Sulphur domains were observed and surface roughness in- both hydrophilic and hydrophobic particles are in contact with water.
creased when the electrochemical potential increased to +258 mV. The interaction of the water molecules with the surface atoms of the
Hampton et al. [60] concluded that the electrochemical oxidation of solid has an influence on the structure of the interfacial water. Knowing
galena is a two-step process: First, sulphur is released into the solution the water structure at the solid-water interface is essential to under-
and then the sulphur deposits on favourable sites; the sulphur domains stand the flotation mechanisms. Currently, the interfacial water struc-
are distributed heterogeneously (Fig. 4), while in Xie's report [57], the ture is attracting great interest from researchers in the colloid and
sulphur domains were more homogeneously distributed. The reason flotation area, and numerous studies have focused on it [69–76].
for this discrepancy may be due to the increased surface roughness Water molecules close to a hydrophilic surface have a strong attractive
of galena in Hampton's experiment. Surface heterogeneity led to a interaction with surface atoms, e.g. via hydrogen bonds, increasing the
heterogeneous distribution of oxidation products. surface free energy. This interaction also leads to a structuring of the
For non-pure minerals which do not have a cleavage face, pre- interfacial water, where water molecules on average reside longer at
treatment such as careful polishing is needed to lower the surface positions with strong interaction than at positions with lower
roughness so that it can be imaged by AFM. Bruening and Cohen [65] interaction force. Such hydration layers typically consist of 3–5 water
applied AFM to identify the surface roughness variations before and molecular layers [69–78].
after coal oxidation. Morga [66] used AFM to study the effect of heat Hydration layers are dynamic structures i.e., the water molecules in
treatment on the surface roughness of semifusinite and fusinite. Heating hydration film are always movable. The water molecules of the first hy-
increased the surface roughness of both semifusinite and fusinite. dration layer are typically completely exchanged within nanoseconds

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.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx 5

Fig. 4. Surface topographies of galena surfaces under different applied potentials: A, +8 mV; B, +258 mV; C, +308 mV; D, +358 mV.
Adapted with permission from [60], Copyright 2011, American Chemical Society.

[79]. At the hydrophobic solid-liquid interfaces, in contrast, a water ex- microscopy (3D-SFM, Fig. 5), the tip is scanned up and down like the
clusion zone with several angstroms was observed using high- needle of a sewing machine while the lateral position is slowly changed
resolution X-ray reflectivity and neutron reflectivity [80–82]. in a scanning motion, which enables to visualize 3D distribution of
To map hydration structures with the tip of AFM is challenging be- water at a mica water interface in a region of 2 × 2 μm in 53 s with an
cause the tip-sample forces induced by the presence of the hydration atomic-scale resolution. In the first 3D-SFM investigation on a mica sur-
layers can be of the same order as the thermal fluctuations in standard face in aqueous buffer solution, a Si cantilever with a resonance fre-
AFM cantilevers. In 2005, Fukuma et al. [83] demonstrated true atomic quency of 123 kHz, a Q factor of 5.8, and 14.2 N/m spring constant in
resolution imaging on a mica surface in aqueous buffer solution by liquid was used. Once complete 3D force field was obtained, any 1D pro-
using an AFM with very low detector noise. The low noise in the system file or 2D cross sections could be extracted. A hexagonal pattern of force
was important because to obtain atomic scale resolution, the amplitude peaks with the same periodicity as the mica substrate was observed.
of the cantilever oscillation had to be on the order of the size of a water Above this first hydration layer, a second and a third layer of force
molecule, i.e. less than 1 nm. Using small amplitude force distance spec- peaks was observed with a lateral shift. Later, similar hydration struc-
troscopy, they also observed oscillations in the tip-sample force which tures were observed on hydrophilic surfaces like calcite, fluorite, and
they ascribed to the layering of water molecules in the tip-sample gap. dolomite [73,78,86]. Most authors so far have used for 3D SFM in fre-
This layering is consistent with earlier reports by Israelachvili and quency modulation, mostly because the force reconstruction from the
Pashley [32] and Zachariah et al. [84]. Israelachvili and Pashley [32] observables frequency and phase shift, and oscillation and drive ampli-
found that the short-range hydration force between two mica surfaces tude are well established [87,88]. Recently, reliable force reconstruction
in water showed an oscillatory profile with a mean periodicity of the methods for amplitude modulation AFM have also become available
water molecule diameter indicating the existence of ordered water [89].
layers. In contrast, Zachariah et al. [84] suggested that the hydrated Although 3D-SFM imaging is always carried out on an atomically flat
ion layering is responsible for the oscillatory hydration force. However, interface, the necessity of this technique becomes more evident when it
information on the potential lateral ordering of water molecules at the is applied to the systems having a larger corrugation and inhomogene-
interface was still missing. Thus, a couple of years later Fukuma ex- ity. Recently, several groups have investigated theoretically by molecu-
tended his high-resolution AFM to enable three dimensional mapping lar dynamics simulations how the measured force field is connected to
of the tip-sample force [85]. In this three-dimensional scanning force the overlap of hydration layers of tip and surface [79,90,91]. It was

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
6 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

Fig. 5. Schematics of (a) 2D-SFM and (b) 3D-SFM. (c) Experimental setup for the developed 3D-SFM [85]. A phase-locked loop (PLL) circuit is used for the resonance frequency shift
detection while an automatic gain control (AGC) circuit is used for keeping the amplitude of 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 Physical Society.

found that the force maxima correspond to the undisturbed positions of microscopy [111], cryo-scanning electron microscopy [112,113], quartz
water molecules at the surface and that the presence of the tip has only crystal microbalance [114], time-resolved fluorescence microscopy
a minor influence on the measured water structure. Further improve- [115], scanning synchrotron based scanning transmission soft X-ray mi-
ments in terms of the measurement technique could be achieved by croscopy [116] and small angle X-ray scattering [117]. The height of
using small cantilevers combining high resonance frequency and low nanobubbles has been reported in the range of 10–100 nm, while the
spring constant [92,93], which provide a better signal-to-noise ratio. length of the three-phase contact line is in general between 50 and
Furthermore, Söngen et al. have improved the 3D scanning method by 500 nm [118]. It is anticipated that nanobubbles should rapidly dissolve
introducing a tip protection mechanism that retracts the tip when a into surrounding water due to a high internal Laplace pressure. How-
threshold force is exceeded at the lower turning point, providing more ever, surprisingly long lifetimes of nanobubbles have been observed
stable imaging [94]. experimentally [95,96]. However, it is out of the scope of this paper to
review the stability theory of nanobubbles; an excellent recent review
is [95].
3.2. Surface nanobubbles characterization Using AFM, the effects of production methods, substrate properties,
salts, solution pH, dissolved gas, surfactants, and temperature on
When hydrophobic mineral surfaces are submerged in water, nanobubbles can be studied. Previous studies showed that degassing
nanobubbles or air are easily formed when the dissolved gas becomes pre-treatment reduces nanobubble formation [119,120]. The gas type
oversaturated for example by increasing temperature or mixing differ- was also found to have great impact on nanobubble nucleation [121].
ent liquids. Such nanobubbles can be stable for many hours [95,96]. Zhang et al. [103,122] found that pH, salts, and sodium dodecyl sulfate
Nanobubbles at solid-water interface are an important subject since has little effect on nanobubbles stability, thus refuting the stability
studies show that an enhanced flotation recovery can be obtained by mechanism of the contamination skin hypothesis. Xu et al. [123]
using nanobubbles [97,98]. Parker et al. [99] first suggested that the dis- observed that the formation of surface nanobubbles was temperature-
continuities in AFM force curves between two hydrophobic surfaces dependent. More nanobubbles were observed at high temperature
were due to the presence of nanobubbles. Experimentally, surface than that at low temperature due to the lower solubility of gas satura-
nanobubbles were firstly imaged by using AFM tapping mode by Ishida tion at high temperature.
et al. [100] and Lou et al. [101]. Subsequently, force mapping mode AFM
[102–104], peak-force tapping mode AFM1 [105–107], and peak-force 4. Reagent adsorption on mineral surface
quantitative nano-mechanics mode AFM2 [108] were all reported as
being capable to image nanobubbles (Fig. 6). In parallel, non-invasive 4.1. Structure of adsorbed reagent
techniques have confirmed the existence of nanobubbles such as
infrared spectroscopy [109,110], total internal reflection fluorescence The interaction between flotation reagents and mineral surfaces are
of particular interest. In flotation practice, various kinds of flotation re-
agents, i.e., collectors, frothers, dispersants, and flocculants, are used to
1
In peak-force tapping the cantilever is periodically moved up and down. In contrast to regulate interfacial properties and thus the interaction between bubble
normal tapping mode, the cantilever is oscillated below resonance frequency. and particle or inter-particle. Here, AFM can help analyzing the adsorp-
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,
tion conformation of surfactants or polymers on mineral surface
adhesion force, and deformation depth can be extracted and quantified from the force- [56,124–131]. A molecular layer or agglomerate can be identified from
distance curve at each pixel. AFM imaging. For example, the adsorption of the surfactants sodium

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx 7

Fig. 6. Nanobubbles pictures imaged using different AFM working modes: (a) tapping mode, Reprinted with permission from [101], Copyright 2000, American Vacuum Society; (b) force
mapping mode, Reprinted with permission from [102], Copyright 2013, American Chemical Society; (c) Peak-force tapping mode, Reprinted with permission from [106], Copyright 2016,
American Chemical Society; (d) Peak-force QNM mode.
Reprinted with permission from [108], Copyright 2013, The Royal Society of Chemistry.

dodecyl sulphate (SDS) and hexadecyltrimethyl ammonium bromide As common practice, guar gum is used as the depressant for molyb-
(CTAB) at graphite–water interfaces has been studied using AFM by denite flotation. The effect of guar gum on molybdenite flotation was
Parachuri et al. [132]. For both surfactants with either positively or neg- studied by Xie et al. [135]. The presence of polymer aggregates on
atively charged head group, the AFM images show the adsorbed surfac- molybdenite surfaces could be correlated with molybdenite flotation re-
tant structures as linear parallel hemi-cylindrical micelles when the sults. The flotation recovery of molybdenite decreased sharply from 69%
concentration is below the CMC concentration. Parachuri et al. [133] fur- to 11% and 3% after guar gum adsorption at concentrations of 1 and
ther studied the effect of cosurfactants (non-ionic 1-dodecanol and cat- 5 ppm. Mierczynska-Vasilev and Beattie [136] studied the adsorption
ionic dodecyl trimethyl ammonium bromide) on the SDS surface of three kinds of substituted carboxymethyl cellulose on talc and
micelle structures. The orderliness of the SDS hemicylinders was re- chalcopyrite. The wettability depressions of three polymers on talc
placed by a herring-bone pattern for the SDS/C12OH system. However, were always more effective than on chalcopyrite since more polymers
in the case of the SDS/C12TAB system, the structures were more similar adsorbed onto talc surface from AFM images. More specifically, low
to the SDS only system. Jaschke et al. [134] used AFM imaged the linear carboxymethyl cellulose coverage on both minerals was observed for
aggregates of ionic surfactant aggregates at a gold surface in aqueous so- a high substitution of the carboxymethyls.
lution. 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 col- 4.2. Single molecule force spectroscopy
lector on a fluorite crystal surface. Both monolayer and bilayer struc-
tures were observed even at low oleate concentration of 10−7 M. Polymers are often used as flocculants in ultra-fine particle flotation
Paiva et al. [129] studied the effect of calcium ions on the adsorption and filtration. Fine particles form large-size flocs under the bridging
of potassium oleate onto apatite surfaces. They found that calcium function of high-molecular-weight polymer. Theoretically, hydrogen
ions play a critical role in potassium oleate adsorption, since the adsorp- bonding, the hydrophobic interaction, electrostatic attraction, and
tion was completed by forming calcium dioleate agglomerates. chemical bonding are the possible driving forces for flocculation. In
Beaussart et al. [56] explored the effect of three kinds of dextrins, a early times it was, however, difficult to measure the force between floc-
regular wheat dextrin (TY), carboxymethyl (CM) dextrin, and hydroxy- culant and mineral particle. With the introduction of single molecule
propyl (HP) dextrin, on molybdenite flotation. Topographies of force spectroscopy (SMFS) it became possible to measure inter- and in-
adsorbed dextrins on molybdenite surface were imaged using AFM tramolecular interaction forces in polymer and supramolecular system
(Fig. 7). The surface coverages of the modified dextrins (CM and HP) with pN resolution using AFM [137–140]. Typically in such experiments,
were much higher than that of regular TY, leading to a lower contact a polymer bridges the AFM probe and substrate. Then the bridge is
angle and flotation recovery. stretched by slowly retracting the tip while measuring the force (Fig. 8).

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
8 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

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.

Measuring the force between polymer flocculant molecule and min- different surface locations, and 2 to 3 substrate-tip pairs were used.
eral surface is of great importance to understand the selective adsorp- N1000 force curves were recorded for each test condition in Long's re-
tion mechanism of the flocculant and thus develop new high port. In deionized water, the adhesion forces were reported to be 40,
efficiency flocculants. Sun et al. [141] carried out SMFS experiments be- 200, and 80 pN on silica, mica, and bitumen surfaces, respectively
tween a novel hybrid polymer Al(OH)3-polyacrylamide (Al-PAM) and a (Fig. 9). When the process water from oil sand processing plant, con-
silica surface. In this case, the Al(OH)3 served as the core connecting a taining various kinds of ions, was used, the forces changed to 50, 100,
number of PAM chains. It was found that the adhesion force between and 40 pN, respectively. The adsorption strength between HPAM and
Al(OH)3 and silica was much higher than that between PAM chain and mica was much higher than that on the bitumen surface. This indicated
silica due to electrostatic attraction. A partially hydrolyzed polyacryl- that a selective flocculation between clay particles, such as mica, can
amide (HPAM) was found to be beneficial to bitumen extraction not achieve when HPAM is added to bitumen-clay mixture suspensions.
only improving bitumen recovery but also accelerating fine solids set- Pensini et al. [143] directly coated carboxymethyl cellulose polymer
tling in the tailings stream. Long et al. [142] measured the adhesion on a fresh mica surface and a silica/borosilicate/iron oxide particle was
force between a partially hydrolyzed polyacrylamide (HPAM) molecule attached to a tip-less cantilever. They recorded retraction forces be-
and different kinds of surfaces (silica, mica, and bitumen). To obtain tween particles and carboxymethyl cellulose coated mica. Different rup-
more representative results for single-molecule desorption, for each ture events could be identified from the shape of the force curves
test condition, the measurement was performed at a number of (Fig. 10). Compared to SMFS experiments, the bridge between the

Fig. 8. (a) Schematic of SMFS; (b) Stretching force curves between a polyprotein molecule and a copper surface. ΔLc is the contour length increment.
Reprinted with permission from [138], Copyright 2012, Nature Publishing Group.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx 9

while a monotonous repulsive force between oxidized coal and paraffin.


Consequently, a fatty acid collector is more suitable for oxidized coal flo-
tation than hydrocarbon oil.

5. Quantification of inter-particle force

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 detrimental effect on flotation
selectivity [148,149]. These slimes follow bubble-particle aggregates
Fig. 9. Adhesion forces between HPAM and different kinds of surfaces in DI water and and enter the froth, lowering the concentrate grade. On the other
commercial plant process water. hand, such slime coatings make targeted particles become hydrophilic
Reprinted with permission from [142], Copyright 2006, American Chemical Society. and prevent bubbles or collectors from adhering to particles, reducing
flotation recovery. Selective flocculation was also widely used for
colloidal probe and mica surface is no longer formed by a single mole- ultra-fine particle flotation to increase bubble-particle collision effi-
cule. Instead a number of molecules interact due to the large contact ciency or reduce gangue minerals entrainment [150]. Measuring the
area. It was found that the adhesion force between iron oxide and driving force responsible for these inter-particle interactions is of both
carboxymethyl cellulose coated mica in Milli-Q water was always academic and practical importance in flotation engineering.
higher than that between silica/borosilicate. Solution pH and ions The quantification of inter-particle force in colloid science is one of
were found to have great impact on the adhesion force. the most important applications of AFM [54]. Early work mainly focused
High molecular-weight flocculants can be attached to the tip and on studying surface forces and verifying the correctness of the
their interaction with a mineral surface is relatively easy to measure. It Derjaguin-Landau-Verwey-Overbeek (DLVO) theory. The critical step
is, however, difficult to directly measure the force between a single sur- in AFM force measurement is to update the sensitivity and calibrate
factant molecule (collector) and a mineral surface. Therefore, some sub- the spring constant of the cantilever. During the data analysis, how to
stitutes were used to model collector-mineral interaction. Fa et al. convert force-displacement curve to force-distance one is also a critical
[144,145] prepared a calcium dioleate micro-sphere to model the inter- step. However, great attention should be paid to the surface roughness
action between the oleate collector and calcite/fluorite. A 10 μm diame- effect since it would influence the force results and thus the repeatabil-
ter calcium dioleate micro-sphere with a surface roughness Ra of 36 nm ity of the experiments. Ducker et al. [52] used AFM to measure the force
for a given 10 μm2 surface area was used. More specifically, strong long- between a hydrophilic silica sphere and a silica plane in presence of so-
range attractive forces in approach curves and adhesion forces in retrac- dium chloride. The measured force at long range could be well predicted
tion force curves were observed between calcium dioleate and fluorite; by the classical DLVO theory. However, a repulsive force at short range
the force between calcium dioleate and calcite was much weaker. In- was observed probably due to the strong hydration of the surfaces. As
deed, such studies indicate that, in the presence of calcium, calcium the DLVO theory treats the intervening medium as continuous, the indi-
dioleate forms firstly in bulk solution and then adsorbs in the form of vidual properties of molecules involved are not taken into consider-
calcium dioleate [146]. ation. Therefore, it cannot describe the hydration force that is required
Xing et al. [147] used solid-state paraffin and stearic acid to repre- to remove the water molecules at the interface when the distance
sent conventional hydrocarbon oil and fatty acid collector in fine coal decreases to a few molecular diameters. Butt [53] also measured the
flotation. A coal particle with diameter of approximately 35 μm was at- force between a silicon nitride tip, alumina, glass, and diamond particle
tached to the end of a tip-less cantilever. The interactions between par- and glass/mica surface in presence of different salt concentrations. A re-
affin/stearic acid and fresh/oxidized coal particles were measured pulsive hydration force with 3 nm decay length between silicon nitride
directly using atomic AFM colloidal probe technique. In this case, the tip and mica was observed at a high salt concentration (N3 M).
paraffin substrate for AFM experiments was prepared using section- Another kind of non-DLVO force is the hydrophobic force. A vast
cutting while stearic acid substrate was prepared by the pellet method. amount of hydrophobic force results can be found in the literature.
More specifically, a significant jump-into contact between oxidized coal Pashley and Israelachvili found that the hydrophobic force is much
particles and stearic acid was observed due to the hydrogen bonding stronger than the van der Waals attraction and a single exponential

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 curves between carboxymethyl cellulose and iron oxide particles in milli-Q water
buffered with NaHCO3 to pH = 8.
Reprinted with permission from [143], Copyright 2013, Elsevier.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
10 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

function gave the best fit to the experimental data [35,36]. Rabinovich 0.001 M KCl solution with varying pH values. Strong long-range attrac-
and Yoon [151] found that the hydrophobic force between a hydropho- tive hydrophobic forces at pH 5.5 and pH 2.7 were observed, confirming
bic glass sphere and a silica plate, could be described with both that hematite is hydrophobic at these pH values. At alkaline pH values,
exponential and power laws. only repulsive forces were found indicating that surface hydroxylation
Up to date, the origin of the hydrophobic force remains unclear. Both takes place at alkaline pH making the hematite surface hydrophilic.
sample preparation methods and experimental techniques have great Often it is more difficult to measure the inter-particle force in the
impact on the result [37]. Nanobubble bridging is one of the most presence of polymer due to experimental difficulties in controlling the
representative mechanisms [152]. Hampton and Nguyen [153,154] polymer dose and the often irreversible adsorption [158,159]. Never-
used a capillary mathematical model to fit the force curves between a theless, Abraham et al. [160] measured the force between silica surfaces
1-octanol esterified silica sphere and a silica plane, which supports in the presence of copolymers of acrylamide and negatively charged
nanobubble bridge hypothesis. Another possible mechanism for the acrylic acid with three different negative charge densities (i.e, 15%,
long-range hydrophobic force is the spontaneous formation of cavities. 40%, and 70% acrylic acid fraction in copolymers). More specifically, at
Cavitation due to metastability of the intervening liquid during ap- low charge density the force was repulsive in the range of experimental
proach or separation was proposed by Christensen and Claesson [155]. polymer concentrations (from 0.1 ppm to 50 ppm). Increasing the
However, detailed discussion about these origin mechanisms is beyond charge density to 40% or 70%, an attractive adhesion force was observed
the scope of this review. Readers who are interested in this topic can at a low polymer concentration due to polymer bridging. A purely repul-
refer to the excellent reviews by Christenson and Claesson [37] and sive force was observed again when the concentration increased to
Meyer et al. [38]. higher levels.
In many cases it is problematic to quantitatively describe inter- Zhou et al. [161] also studied the effect of charge density (10, 40, and
particle forces because of irregular particle structure and surface hetero- 100%) of cationic polymers on the force between silica surfaces using
geneity of practical minerals. Still, the force curves can be qualitatively AFM. The 10% charged polymer produced steric repulsion upon ap-
compared with varying the concentration of salt or pH and may provide proach and long-range adhesion force at the optimum flocculation con-
guidance for flotation engineering. Xing et al. [148,149] studied the ef- centration. In contrast, the 40 and 100% charged polymers produced
fect of calcium ions on coal flotation in the presence of kaolinite clay. attraction during the approach, due to the attractive electrostatic
The interaction force between coal and kaolinite was measured using force, and strong adhesion during the retraction stage at optimum poly-
AFM (Fig. 11). No jump-in was observed in de-ionized water and the mer dosages. It was also found that the polymer dose that produced the
repulsive force corresponded to that in the case of no kaolinite coating. optimum flocculation and the maximum compressive yield stress typi-
A sudden jump-in was found when Ca2+ was added to a concentration cally corresponded to the polymer concentration that produced the
of 3 mM. The repulsive electrostatic force was suppressed by excessive maximum adhesion force in AFM force measurement for each polymer.
Ca2+ addition, and the attractive force began to dominate the Ofori et al. [162] examined the coal tailings flocculation performance
kaolinite-particle interaction. Therefore, the presence of Ca2+ in practical using different types and concentrations of flocculants. Averaged jump-
flotation aggravates slime coating and deteriorates flotation selectivity. out distances were statistically determined from thousands of AFM
The effect of clay types, i.e., kaolinite and montmorillonite on fine force curves. The optimal flocculant concentration would be the point
coal flotation was further studied [156]. Only a short jump-into- corresponding to the maximum jump-out distance in AFM force curves.
contact was observed in de-ionized water for coal-montmorillonite. However, flocculation and sedimentation tests showed optimum
This low weak attraction correlates with a lower flotation recovery flocculant concentrations were much higher than those corresponding
and worse selectivity for the coal-montmorillonite system compared to the optimum adhesive strength found in the AFM measurements.
with that of coal-kaolinite system where a repulsive force dominated. This was probably due to the large surface area of particles in the con-
Shrimali et al. [157] prepared two kinds of hematite colloidal probes centrated suspension used in the settling tests compared with just a
of around 15 μm in diameter. In one case, the hematite was glued in a few particles used in the atomic force measurement. In concentrated
way that the (001) crystal face was exposed, while in the other case suspension, so much polymer adsorbs to the particle surface that
the hematite particle was oriented so that the (100) crystal surface polymer is depleted. This would also lead to less polymer at the particle
was exposed. AFM force measurements between the (001) hematite surfaces. It should be noted that this effect is only important at low
crystal surface and the above two colloidal probe were carried out in polymer concentrations.

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

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx 11

Recently, a multi-particle colloidal probe technique (MPCPT) based form of CaOH+ and MgOH+. Thus, these monohydroxyl species act as
on AFM-inverted optical microscope has been developed to measure bridges between sodium dodecylbenzene sulfonate and the solid
inter-particle forces in situ [159,163]. MPCPT works in a colloidal surface.
suspension in presence of a number of particles, which simplifies the Liu et al. [169] measured the force between bitumen and different
precise dosing of polymers and makes this approach less sensitive to kinds of irregular clay particles (montmorillonite and kaolinite). When
impurities. Note that still one particle was attached to the end of canti- using irregular particles as a colloidal probe a quantitative comparison
lever in MPCPT. In these experiments, silane reagents are used to of measured forces with the standard DLVO theory of colloid stability
hydrophobize a glass substrate and a tip-less AFM cantilever. A number is difficult because the effectively interacting areas and the contact
of polystyrene latex particles are injected into the liquid cell. The parti- area are unknown. Consequently, the force curves have to be only
cles deposit onto glass surface. Then, a particle is picked up by moving qualitatively compared. In Liu's experiments [169] the adhesion force
the cantilever on its upside. Finally, force measurement is conducted be- between bitumen and montmorillonite was always higher than that be-
tween this colloid probe and another deposited particle. The main ad- tween bitumen and kaolinite, especially in in the presence of Ca2+. This
vantage of MPCPT, compared with that of traditional colloid particle is probably due to the large capacity of adsorption of calcium for
probe only with one or two colloidal particles, is that the adsorption of montmorillonite, depressing the repulsive force barriers and enhancing
polyelectrolyte on particles can be controlled precisely, due to the larger the adhesion force.
internal surface area in multi-particle suspension system. A related subject which should be briefly mentioned is bacteria-
Flotation is also used for oil sands separation engineering. Oil sand mineral interaction. A deep understanding of bacteria-mineral interac-
separation has been studied, e.g. by Xu's and his team from Edmonton tion force is critical to mineral bioleaching engineering. To measure
[164]. In oil sand processing, the first step is to liberate oil from the the force between individual bacteria and a mineral surface by AFM,
sand. The sand typically consists of carbonate rocks and silica. Air bub- the main challenge is to anchor the bacteria onto the cantilever or flat
bles are introduced to capture the heavy oil (bitumen) leaving sand par- substrate without affecting cell activity [170–172].
ticles in the pulp. Clays account for 15–30% in raw oil sands [165]. Lower et al. [173] introduced a new technique named biologically-
Important sub-processes are: bitumen-sand liberation, bitumen- active-force probe (BAFP) to measure the force between bacteria and
bubble interaction, and bitumen-fine particle interaction (clay coating). different mineral surfaces (muscovite, goethite, and graphite). They
Here, we treat bitumen as the targeted particle similar to mineral flota- used poly-D-lysine to functionalize glass beads; the negatively charged
tion; therefore, we also briefly discuss the application of AFM in glass and cell-surfaces were bridged by positive charged poly-D-lysine
bitumen-sand liberation and clay coating in this section. in situ (Fig. 13). It was found that ions, mineral surface charge, and
Liu et al. [166] coated a thin bitumen film on a flat silica wafer with a hydrophobicity significantly affected the interaction. Diao et al. [174]
spin coater and then measured the force between bitumen and silica compared adhesion forces between chalcopyrite and acidithiobacillus
particle in 1 mM KCl solution. They found that both pH and calcium con- thiooxidans/acidithiobacillus ferrooxidans under various pH conditions.
centration play an important role in bitumen-silica interaction (Fig. 12). The average adhesion force for acidithiobacillus ferrooxidans was
Not only did the repulsive barrier increase, but also the adhesion force always stronger than that of acidithiobacillus thiooxidans, especially at
decreased with increasing pH or decreasing calcium concentration. low pH.
Temperature is an important parameter in oil sand processing.
Increasing temperature reduces bitumen viscosity and thus decreases 6. Bubble-particle interaction and thin liquid film drainage
the adhesion force [167]. Alkaline solution with high temperature
and low ion concentrations promotes bitumen liberation. Zhang et al. Bubble-particle interaction is the critical step for flotation. It is well
[168] further suggested a synergistic effect between surfactants accepted that the hydrophobic force is responsible for successful
(dodecyltrimethylammonium chloride and sodium dodecylbenzene bubble-particle attachment [42]. AFM is a unique tool to directly
sulfonate) and divalent cation ions (Ca2+ or Mg2+) for bitumen– measure bubble-particle interaction. Note that AFM force measurement
silica interaction. More specifically, in alkaline solution, sodium between an air bubble and a solid particle is accompanied with the
dodecylbenzene sulfonate does not adsorb on negatively charged silica deformation of gas/liquid interface under both the hydrodynamic and
and bitumen surfaces due to the strong repulsive double-layer force. surface forces. The dynamic coupling between force, bubble deforma-
However, this is not the case when divalent cation ions were added. tion and film drainage makes both theoretical analysis and experimen-
Ca2+ or Mg2+ prefers adsorbing on silica and bitumen surface in the tal verification challenging [41,175].

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
concentration. The dashed and solid lines represent the experimental and theoretical fitting results, respectively.
Reprinted with permission from [166], Copyright 2005, Elsevier.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
12 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

Fig. 13. (A) Schematic of BAFP, (B) Scanning laser confocal image of BAFP. The green particle is the cell coated glass. Scale bar is 10 μm. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)
Reprinted with permission from [173], Copyright 2005, Elsevier.

Ducker et al. [176] and Butt [177] first used AFM to measure the attractive force increased whereas the wetting film rupture thickness
force between a bubble and a particle in 1994. A small hydrophilic or hy- did not changed significantly. Preuss and Butt [195,196] studied the in-
drophobic silica particle was attached to the end of AFM cantilever and fluence of dodecyltrimethylammonium bromide (DTAB) and sodium
then approached to the bubble surface at a low speed in pure water. The dodecyl sulfate (SDS) on bubble-silica interaction. When DTAB was
observed attractive force between a hydrophilic particle and an air added at concentrations below the CMC a hydrophobic attraction was
bubble in Ducker's experiment [176] probably due to contamination. observed caused by the strong adsorption of trimethylammonium
When surfactant was added, the long range attractive force disappeared head group to the silica. The exposed dodecyl group rendered the
and a monotonous repulsive force emerged. In Butt's report [177], he particles hydrophobic. When the DTAB concentration exceeded 6 mM,
found that a repulsive force acted between bubble and a hydrophilic the repulsive force dominated bubble-silica interaction again. Above
glass. In contrast, a jump-into contact was observed when hydrophobic 6 mM DTAB adsorbs as a double layer to silica exposing the positively
particles approached the bubble surface and a three-phase contact was charged head group to the water phase. The air water interface is also
formed. In general, particles with finite contact angle towards water positively charged due to adsorption of DTAB, leading to a strong
form a three-phase contact and attach to the air/water interface. electrostatic repulsion. In the experimental concentration range of
The determination of zero distance is an essential step to convert SDS, no attractive force was observed between hydrophilic silica and
AFM force-displacement curve to force-distance curve [178–181]. For bubble. For a hydrophobic particle, hydrophobic force dominated the
the interaction between solid surfaces, the zero distance is deduced total force. However, the range and magnitude of the attractive force
from constant compliance regime in a hard contact system. For the decreased as SDS concentration increased and SDS adsorbs to the
interaction of a particle with an air/water interface zero distance is a hydrophobic solid surface and the air/water interface.
matter of definition. It is convenient to set zero distance to the Nguyen et al. [180–182] did a series of experiments on bubble-
equilibrium position of a particle in the air/water interface. The bubble particle interaction using AFM. For a hydrophilic silica particle, the
is assumed to behave as a Hookean spring under external force forces at different electrolyte concentrations could be well fitted by
[180,182–186]. To illustrate the situation we consider a small spherical using classical DLVO theory. With increasing ion concentration, the
particle at a planar air/water interface as a model. For particles much electrostatic force decreased due to the compression of the double
pffiffiffiffiffiffiffiffiffiffiffi
smaller than the capillary length κ ¼ γ=ρg capillary forces dominate; electrostatic layer. The effect of approach velocity on the force between
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 γ = 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]:

2πγδ
F¼ ð1Þ
0:809 þ ln ðκ=RÞ

Eq. (1) has been derived on theoretical grounds. It agrees with


experimental results, e.g. for the interaction of a microparticle with a
bubble [176,181], for a microparticle moved against the surface of
water from the water side [192,193], and for particles being drawn
out of a liquid-air interface [194]. When the contact line is pinned,
Eq. (1) is also valid, only R has to be replaced by Rsinβ, where β denotes
the position of the pinned contact line (Fig. 14). Fig. 14. Schematic of a small particle at a liquid surface in equilibrium (bottom) and when
Ishida [183] studied the effect of surface hydrophobicity on bubble- a force is pulling the particle upwards (top). The contact line, characterized by the angle β,
particle interaction force. With increasing contact angle, the range of the is assumed to slide over the particle surface at constant contact angle Θ.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx 13

a hydrophilic glass sphere and bubble was also studied [180]. The decay length of 0.8 nm. Increasing the contact angle to 85°, both the crit-
repulsive hydrodynamic force increased monotonically as the ap- ical film rupture thickness and hydrophobic force decay length in-
proach speed increased. When piezo approach speeds were lower creased, illustrating that the hydrophobic force increased with water
than 0.6 μm/s, the hydrodynamic force was negligible and instead contact angle. Force measurements between air bubbles and sphalerite
surface forces were dominant. When the particle became hydrophobic, and molybdenite surfaces of different water contact angle were also
a sudden jump-into contact due to the attractive hydrophobic force al- conducted by Xie et al. [135,201]. More specifically, the effect of poly-
ways observed at a low approach speed. The attractive hydrophobic mer depressant (guar gum) on the bubble-molybdenite attachment
force also increased as water contact angle increased. Nguyen et al. was studied. When 1 ppm guar gum was added, polymer aggregates
[197] used AFM to determine the contact angle of a micrometer-sized on the surface became visible from AFM images, resulting in a decreased
polyethylene sphere according to the depth of the particle penetration bubble-particle attachment. Further increasing guar gum concentration
into the bubble at the zero force. They observed that the contact angle to 5 ppm, the contact angle of molybdenite surface decreased from
measured by AFM changes with the speed of the AFM piezo. ≈74° to 65°. A monotonic repulsive force was observed. The AFM
Ishida [183] explored the particle hydrophobicity on the force be- force results were consistent with the practical molybdenite flotation
tween a silica particle and an air bubble. It was found that the hydropho- results. The final flotation recovery decreased from 69% to 11% and to
bicity of the particle did not significantly change the film rupture 3% after guar gum adsorption at 1 and 5 ppm concentration solutions.
thickness, whereas the pH of the solution played a critical role in critical The other effective approach complementary to AFM force measure-
film rupture thickness. Assemi et al. [181] used AFM to measure the ment is optical interferometry. Clark et al. [202–205] first described the
force between a silica and an air bubble in deionized water and KCl so- use of evanescent wave scattering to measure the separation between a
lutions with a low piezo speed. By fitting the force curves using DLVO solid surface and a particle that was attached to an AFM cantilever. The
theory, the surface potential on the bubble surface was obtained. The evanescent scattering apparatus was essentially identical to that used in
force curves at the edge of the bubble could fit very well to theory, the original total internal reflection microscopy except that the light
even at high applied forces due to the less bubble deformation effect. scattered back into the incident medium was collected. However, for
However, the simple assumption that the bubble surface behaves as bubble-particle interaction, the evanescent wave scattering is not suit-
a linear spring will not always be appropriate [175,198]. Chan [175] able because the liquid has a higher refractive index than air. In 2015,
derived a no-linear force-displacement relation from the solution Shi et al. [206] first combined a reflection interference contrast micro-
of the Young-Laplace equation between a bubble probe and a solid scope (RICM) with an AFM to measure the force and the thin water
plane, as shown in Eq. (2): film profile between a bubble and mica surface simultaneously
        (Fig. 15). A bubble was attached on the end of a cantilever. The dynamic
F F 1 1 þ cosθ 1 profiles of air-water interface were reconstructed using RICM with
ΔD ¼ log þ 2 1 þ log − −1
4πγ 8πγRb 2 1−cosθ 2 þ cosθ nanometer resolution in normal direction. It was found that the
ð2Þ Stokes-Reynolds-Young-Laplace model can predict both the force and
film drainage with a no-slip boundary condition. A stable wetting film
The additional logarithmic term leads to a weak deviation from a was observed between an air bubble and a hydrophilic mica surface.
linear-distance relationship when the compression ΔD is not negligible In contrast, jump-in events where the interaction force drastically
compared to the radius of the bubble. turned from repulsive to attractive were observed for hydrophobized
In recent years, significant progress has been achieved on both mica with both 45° and 90° contact angle (Fig. 16). Again, the critical
hydrodynamic drainage modelling and AFM-based experimental film rupture thickness increased as contact angle increased. Due to the
techniques. It has become possible to measure the force, bubble short range of van der Waals force and the suppression of the electro-
deformation, and liquid film drainage simultaneously [41]. static force at 500 mM NaCl, an additional attractive hydrophobic
Chan et al. [175,199] derived a model called augmented Stokes- force was considered as the reason for triggering film rupture. From
Reynolds-Young-Laplace model to predict the force and the evolution the Stokes-Reynolds-Young-Laplace model, the effective hydrophobic
of the wetting film profile before the formation of three-phase contact decay length was inferred to be 0.8 and 1.0 nm for the 45° and 90°
line in dynamic AFM experiments. The defining equations of the water contact angle, respectively. These results are explained in terms
augmented Stokes-Reynolds-Young-Laplace model under a no-slip
boundary condition are [175]:
 
∂h 1 ∂ 3 ∂p
¼ rh ð3Þ
∂t 12μr ∂r ∂r
 
2γ γ ∂ ∂h
p¼ −Π− r ð4Þ
Rb r ∂r ∂r

Eqs. (3) and (4) can be solved numerically in the region 0≤r ≤ rmax
when h b bRb. rmax is selected as the value where the local separation
h is so large that the contribution of disjoining pressure could be
neglected. The Stokes-Reynolds-Young-Laplace model has the ability
to capture the essential physical characteristics in AFM force measure-
ments and provides a basic understanding of bubble-particle interaction
force and film drainage dynamics.
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 force curves were fitted
by using the Stokes-Reynolds-Young-Laplace model. A jump-into
contact at 7.5 nm was experimentally observed for a partially Fig. 15. Schematic of the AFM-RICM experimental setup.
hydrophobized mica with 45° water contact angle. The model Adapted with permission from [206], Copyright 2015, American Chemical
reproduced the force results assuming a hydrophobic force with a Society.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
14 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

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.

of a different water structure at the solid-water interface with different fitted with classical DLVO theory. They found that pH has little effect
surface contact angle. on the surface potential of muscovite basal face. However, on the edge
Combining AFM with RICM is a landmark in our full understanding they could not get good results due to the roughness of the mica edge
of bubble-particle interaction [207]. In the next few years, synchronous surface over the large contact area.
measurement of the force and the spatiotemporal evolution of thin To overcome the surface roughness effect, Gupta and Miller [62] di-
water film profile between bubble and particle in presence of flotation rectly used a pyramid-shaped silicon nitride tip as the probe to map the
reagents will shed new light on flotation mechanism [41]. surface potential of two kaolinite basal faces (silica face and alumina
face). The silicon tip was modeled as a cone with a spherical cap at its
7. Other applications apex. Then, the DLVO forces between tip and flat surface could be rea-
sonably approximated as the sum of the conical-substrate interaction
In recent years, several studies aimed to determine the mineral sur- force and the cap-substrate interaction force [214]. The isoelectric
face charge and wetting characteristics using direct AFM force measure- point of the silica tetrahedral face was lower than that of the alumina
ment [208]. Historically, a colloid probe has usually been employed and octahedral face. Yan et al. [215] further used this approach to explore
the surface charge information was obtained by fitting the force curves the surface potentials of both basal and edge faces of talc and muscovite.
with the DLVO theory. Ducker et al. [52] applied AFM to measure the The ultramicrotome cutting technique was applied to prepare the edge
force between silica surfaces at different salt concentrations and pH face. The surface potentials of basal faces of both talc and muscovite
values. The force curves were fitted using DLVO theory, where the were always more negatively charged than that of edge faces in the
surface potential of silica was used as the fitting parameter. The similar range of experimental pH (5.6–10.1) (Fig. 17). This is due to the differ-
procedure was also adopted to determine the surface potentials of mica/ ent charging mechanisms between basal and edge faces. The permanent
silica and the isoelectric point of fluorite by Hartley et al. [209] and negative charge of basal faces is attributed to fixed isomorphic substitu-
Assemi et al. [210]. tion while protonation-deprotonation reactions are responsible for the
Another aspect is to measure the point of zero charge (PZC) of pH-dependent properties of the edge face. The anisotropic surface po-
different faces of anisotropic layered silicates such as kaolinite, talc, tentials of scheelite crystal and molybdenite were also studied using
pyrophyllite and illite. Traditional zeta potential measurements by AFM force measurements [58,216]. The 101 face of scheelite crystal
electrophoresis or streaming potential assume a uniform charge density was the most negatively charged surface, followed by 112 and 001
on these layered silicate particles. Often, however, macroscopic coagula- faces. For molybdenite, it was found that the surface potentials of both
tion of dispersed particles is not necessarily observed at the PZC. The edge and basal faces are highly pH-dependent.
reason for coagulation lies in the fact that the surface charge is not dis- One can use a hydrophobized probe to map the surface hydropho-
tributed evenly and that on one particle regions with positive and neg- bicity by fitting the force profile using extended DLVO theory with an
ative charge density may exist. However, the problem can be solved additional hydrophobic force taken into consideration. A high attractive
using AFM since with the AFM local charge densities can be measured force corresponds to high surface hydrophobicity (large local water
[211,212]. Zhao et al. [213] used an ultramicrotome cutting technique contact angle). Lu et al. [58] used a single exponential function to fit
to prepare an edge surface of muscovite, while a smooth basal face the experimental hydrophobic force curve between an OTS-coated tip
was easily prepared by cleavage. A silica sphere with an 8 μm diameter and a molybdenite surface in 10 mM NaCl solution at various pH
was attached to the cantilever and a series of force curves were ob- (Fig. 18). An attractive hydrophobic force at all pH ranges was observed.
tained. The force profiles between the silica and basal face were well However, this was not the case between the tip and edge surface. This

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx 15

Fig. 17. Force curves between AFM tip and different faces in 1 mM KCl solution: A, talc basal plane; B, talc edge plane; C, muscovite basal plane; D, muscovite edge plane. Symbols represent
experimental data and the solid lines represent theoretical fits.
Adapted with permission from [215], Copyright 2011, American Chemical Society.

illustrated that the basal surface of molybdenite shows a hydrophobic CuSO4 activation and amyl xanthate conditioning. Amyl xanthate ad-
character, while the edge surface is hydrophilic. This observation is con- sorption increased surface hydrophobicity.
sistent with the contact angle results. Xie et al. [217] adopted the same Colloidal probe AFM can also be applied to characterize hydrophobic
approach to map the sphalerite surface hydrophobicity before and after surface heterogeneities on a micro scale. The colloidal probe-sample

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

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
16 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

The main drawback of AFM is the absence of a chemical imaging ca-


pability. It is necessary to develop new AFM-based techniques that have
the ability to analyze both surface morphology and chemical composi-
tion in the contact area. Optical spectroscopy analysis will be the opti-
mal choice. Recently, significant progress has been achieved on the
hybrid AFM techniques such as the combination of AFM and infrared
spectroscopy (IR) and Raman spectroscopy, which show great poten-
tials in flotation research [221–223]. For AFM-IR, the 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 limi-
tation (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 enhance-
Fig. 19. Schematic diagram of AFM-IR. IR laser is focused on a sample near the tip of an ment of the electromagnetic field at the tip end can be achieved. The
AFM. The AFM tip is used as a local detector of IR absorption.
IR absorption of the substrate will cause a local thermal expansion and
Adapted with permission from [221], Copyright 2016, American Chemical Society.
an additional dipolar interaction between tip and sample. By measuring
the response of the cantilever to IR absorption as a function of wave-
interaction [218] and parameters derived from force-distance curves length, IR absorption spectra of a nano-scale region are obtained. The
[219] in the context of flotation or other processes like filtration can IR laser can be also fixed at a constant wavelength and IR absorption
be obtained. Rudolph and Peuker [220] combined hydrophobic colloidal as a function of position across the sample to create chemical images
probe AFM with Raman spectroscopy to evaluate the wettability of min- that show the distribution of chemical species [221]. AFM-IR is typically
erals in finely inter-grown ore in situ. The type of mineral phase, where performed by detecting the oscillation amplitude of the cantilever as a
force measurements were conduct, could be identified by Raman mi- function of sample position or wavelength. Detailed information on
croscopy. They proposed that adhesion force mapping is a more precise AFM-IR can be seen in [221]. Fig. 20 shows a resonance-enhanced
indicator for the floatability of practical minerals. The combined colloi- AFM-IR image of a self-assembled monolayer of alkyl thiol ethoxylate
dal probe AFM and Raman technique will also be a powerful tool to in- (PEG) on gold surface. The IR absorption band at 1340 cm−1 was attrib-
vestigate mineral-reagent interactions in flotation systems. uted to a CH2 peak, and the AFM-IR image directly confirmed the loca-
tion of the PEG island regions.
Raman spectroscopy, complementary to IR, is an alternative tool for
8. Conclusions and perspectives identifying chemical species. The combination of AFM and Raman also
offers the opportunity to obtain both surface morphology and chemical
AFM is an established tool in mineral flotation used both for imaging composition of the sample [222]. In this context, more complementary
and a force sensing. It can be successfully used to image the minerals, to information on mineral surface/mineral-water interface can be ob-
characterize the water structure at mineral-water interface, to study the tained. For example, the specific adsorption of flotation reagents on
adsorption of flotation reagents on mineral surfaces, to measure inter- mineral surface can be identified successfully. Note that the chemical
particle forces directly, and characterize bubble-particle interaction. De- image obtained by AFM-IR is always affected by surface characteristic
tailed information at the nanoscale has been obtained, greatly assisting such as roughness. Tip-enhanced IR spectroscopy is difficult to operate
the fundamental understanding of flotation. One challenge is to mea- in aqueous medium due to the strong IR absorption bands of water.
sure precise shapes of water/air interfaces in particle-bubble experi- By contrast, AFM-Raman may allow the experiments conducted in
ments due to the absence of hard contact point. aqueous environments, while the reproducibility of the Raman signal

Fig. 20. (a) AFM topographic image of a self-assembled PEG monolayer on gold (b) AFM-IR absorption image at 1340 cm−1 peak, corresponding to the CH2 peak as shown in the (c) AFM-IR
spectrum of the monolayer. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
Image reprinted with permission, Copyright 2014, Anasys Instruments.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx 17

enhancement is still a problem. Nevertheless, the complementary char- [16] Vashisth S, Bennington CPJ, Grace JR, Kerekes RJ. Column flotation deinking: state-
of-the-art and opportunities. Resources, Conservation and Recycling 2011;55:
acterization of chemical composition using optical spectroscopy for 1154–77.
AFM topography imaging and the synchronous measurement of the [17] Hosseini S, Forssberg E. XPS & FTIR study of adsorption characteristics using cat-
force and distance involving deformable bubble as a force sensor will ionic and anionic collectors on smithsonite. Journal of Minerals and Materials Char-
acterization and Engineering 2006;5(1):21–45.
become an active area over the coming years and further shed new [18] Beattie DA, Kempson IM, Fan LJ, Skinner WM. Synchrotron XPS studies of collector
light on flotation mechanism. adsorption and co-adsorption on gold and gold: silver alloy surfaces. Int J Miner
Process 2009;92(3):162–8.
[19] Giesekke EW. A review of spectroscopic techniques applied to the study of interac-
Symbols tions between minerals and reagents in flotation systems. Int J Miner Process 1983;
11(1):19–56.
[20] Hosseinpour S, Tang F, Wang F, Livingstone RA, Schlegel SJ, Ohto T, et al.
F force Chemisorbed and physisorbed water at the TiO2/water interface. Journal of
g acceleration of gravity Physical Chemistry Letters 2017;8(10):2195–9.
[21] Schaefer J, Gonella G, Bonn M, EHG Backus. Surface-specific vibrational spectros-
h thickness of the intervening water film or separation distance
copy of the water/silica interface: screening and interference. Phys Chem Chem
p hydrodynamic pressure difference Phys 2017;19(25):16875–80.
R radius of particle [22] Xu Z, Liu J, Choung J, Zhou Z. Electrokinetic study of clay interactions with coal in
flotation. Int J Miner Process 2003;68:183–96.
Rb radius of bubble before deformation
[23] Liu J, Zhou Z, Xu Z, Masliyah J. Bitumen–clay interactions in aqueous media studied
r radial distance from the center of the film by zeta potential distribution measurement. J Colloid Interface Sci 2002;252:
δ the distance the particle moved from its equilibrium position 409–18.
ρ density [24] Wu C, Wang L, Harbottle D, Masliyah J, Xu Z. Studying bubble–particle interactions
by zeta potential distribution analysis. J Colloid Interface Sci 2015;449:399–408.
κ capillary length [25] Ralston J, Fornasiero D, Hayes R. Bubble–particle attachment and detachment in
ΔD AFM piezo displacement flotation. Int J Miner Process 1999;56:133–64.
γ surface tension [26] Nguyen A, Evans G, Schulze H. Prediction of van der Waals interaction in bubble–
particle attachment in flotation. Int J Miner Process 2001;61:155–69.
θ water contact angle [27] Dai Z, Fornasiero D, Ralston J. Particle–bubble collision models—a review. Adv
μ viscosity Colloid Interface Sci 2000;85:231–56.
Π disjoining pressure [28] Koh P, Schwarz M. CFD modelling of bubble–particle collision rates and efficiencies
in a flotation cell. Miner Eng 2003;16:1055–9.
[29] Chau TT, Bruckard WJ, Koh P, Nguyen AV. A review of factors that affect contact
angle and implications for flotation practice. Adv Colloid Interface Sci 2009;150
Acknowledgments (2):106–15.
[30] Butt HJ, Kappl M. Surface and interfacial forces. WILEY-VCH: Weinheim; 2010.
This research was supported by the National Natural Science Foun- [31] Israelachvili JN. Intermolecular and surface forces. Third Edition. USA: Academic
press; 2011.
dation of China (grant no. 51774286, 51574236, 51574240), a project [32] Israelachvili JN, Pashley RM. Molecular layering of water at surfaces and origin of
funded by the China Postdoctoral Science Foundation (2015T80606, repulsive hydration forces. Nature 1983;306:249–250.33.
2014M550317) for which the authors express their appreciation. [33] Butt HJ. Measuring electrostatic, van der Waals, and hydration forces in electrolyte
solutions with an atomic force microscope. Biophys J 1991;60:1438–44.
Yaowen Xing also appreciates China Scholarship Council for the finan-
[34] Israelachvili J, Wennerström H. Role of hydration and water structure in biological
cial support for their research stay at Max Planck Institute for Polymer and colloidal interactions. Nature 1996;379(6562):219–25.
Research. HJB also acknowledges support by the Max Planck Center [35] Israelachvili JN, Pashley RM. The hydrophobic interaction is long-range, decaying
for Complex Fluid Dynamics Fluid Dynamics of Complexity. exponentially with distance. Nature 1982;300:341–2.
[36] Israelachvili JN, Pashley RM. Measurement of the hydrophobic interaction between
two hydrophobic surfaces in aqueous electrolyte solutions. J Colloid Interface Sci
References 1984;98:500–14.
[37] Christenson HK, Direct Claesson PM. Measurements of the force between hydro-
[1] Nguyen AV, Schulze HJ. Colloidal science of flotation. Marcel Dekker Inc: New York, phobic surfaces in water. Adv Colloid Interface Sci 2001;91:391–436.
America; 2004. [38] Meyer EE, Rosenberg KJ, Israelachvili J. Recent progress in understanding hydro-
[2] Ejtemaei M, Gharabaghi M, Irannajad MA. Review of zinc oxide mineral beneficia- phobic interactions. Proc Natl Acad Sci 2006;103:15739–46.
tion using flotation method. Adv Colloid Interface Sci 2014;206:68–78. [39] Laskowski JK, Kitchener JA. The hydrophilic-hydrophobic transition on silica. J Col-
[3] Aghazadeh S, Mousavinezhad SK, Chemical Gharabaghi M. Colloidal aspects of loid Interface Sci 1969;29:670–9.
collectorless flotation behavior of sulfide and non-sulfide minerals. Adv Colloid In- [40] Blake TD, Kitchener JA. Stability of aqueous films on hydrophobic methylated silica.
terface Sci 2015;225:203–17. Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Con-
[4] Wu ZJ, Wang XM, Liu HN, Zhang HF, Miller JD. Some physicochemical aspects of densed Phases 1972;68:1435–42.
water-soluble mineral flotation. Adv Colloid Interface Sci 2016;235:190–200. [41] Xing Y, Gui X, Pan L, Pinchasik B-E, Cao Y, Liu J, et al. Recent experimental advances
[5] Xing YW, Gui XH, Liu JT, Cao YJ, Lu Y. Effects of energy input on the laboratory col- for understanding bubble-particle attachment in flotation. Adv Colloid Interface Sci
umn flotation of fine coal. Separation science. Dent Tech 2015;50:2559–67. 2017;246:105–32.
[6] Xing YW, Gui XH, Cao YJ, Wang DP, Zhang HJ. Clean low-rank-coal purification [42] Xing Y, Gui X, Cao Y. The hydrophobic force for bubble-particle attachment in
technique combining cyclonic-static microbubble flotation column with collector flotation-a brief review. Phys Chem Chem Phys 2017. https://doi.org/10.1039/
emulsification. J Clean Prod 2017;153:657–72. C7CP03856A.
[7] Xing YW, Gui XH, Cao YJ, Wang YW, Xu MD, Wang DY, et al. Effect of compound [43] Binnig G, Quate CF, Gerber C. Atomic force microscope. Phys Rev Lett 1986;56:
collector and blending frother on froth stability and flotation performance of oxi- 930–3.
dized coal. Powder Technol 2017;305:166–73. [44] Drake B, Prater CB, Weisenhorn AL, Gould SAC, Albrecht TR, Quate CF, et al. Crystals,
[8] Rubio J, Souza ML, Smith RW. Overview of flotation as a wastewater treatment polymers, and processes in water with the atomic force microscope. Science 1989;
technique. Miner Eng 2002;15:139–55. 243:1586–9.
[9] Cai X, Chen J, Liu M, Ji Y, Numerical An S. Studies on dynamic characteristics of oil- [45] Butt HJ, Berger R, Bonaccurso E, Chen Y, Wang J. Impact of atomic force microscopy
water separation in loop flotation column using a population balance model. Sep on interface and colloid science. Adv Colloid Interface Sci 2007;133:91–104.
Purif Technol 2017;176:134–44. [46] Rugar D, Hansma P. Atomic force microscopy. Physics Today 1990;43(10):23–30.
[10] Rao F, Liu Q. Froth treatment in athabasca oil sands bitumen recovery process: a [47] Meyer E. Atomic force microscopy. Progress in Surface Science 1992;41(1):3–49.
review. Energy Fuel 2013;27:7199–207. [48] Jalili N, Laxminarayana K. A review of atomic force microscopy imaging systems:
[11] Kasongo T, Zhou Z, Xu ZH, Masliyah J. Effect of clays and calcium ions on bitumen application to molecular metrology and biological sciences. Mechatronics 2004;
extraction from Athabasca oil sands using flotation. Can J Chem Eng 2000;78: 14(8):907–45.
674–81. [49] Maver U, Velnar T, Gaberšček M, Planinšek O, Finšgar M. Recent progressive use of
[12] Zhou F, Yan C, Wang H, Zhou S, Liang H. The result of surfactants on froth flotation atomic force microscopy in biomedical applications. TrAC Trends Anal Chem 2016;
of unburned carbon from coal fly ash. Fuel 2017;190:182–8. 80:96–111.
[13] Altun NE, Xiao CF, Hwang JY. Separation of unburned carbon from fly ash using a [50] Albrecht TR, Grutter P, Horne D, Rugar D. Frequency modulation detection using
concurrent flotation column. Fuel Process Technol 2009;90:1464–70. high-Q cantilevers for enhanced force microscope sensitivity. J Appl Phys 1991;
[14] Bartonova L. Unburned carbon from coal combustion ash: an overview. Fuel 69(2):668–73.
Process Technol 2015;134:136–58. [51] Jarvis SP, Ishida T, Uchihashi T, Nakayama Y, Tokumoto H. Frequency modulation
[15] Finch JA, Hardie CA. An example of innovation from the waste management detection atomic force microscopy in the liquid environment. Applied Physics A
industry: deinking flotation cells. Miner Eng 1999;12:467–75. 2001;72(1):S129–32.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
18 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

[52] Ducker WA, Senden TJ, Pashley RM. Direct measurement of colloidal forces using [88] Welker J, Illek E, Giessibl FJ. Analysis of force-deconvolution methods in frequency-
an atomic force microscope. Nature 1991;353:239. modulation atomic force microscopy. Beilstein Journal of Nanotechnology 2012;3:
[53] Butt HJ. Measuring electrostatic, van der Waals, and hydration forces in electrolyte 238–48.
solutions with an atomic force microscope. Biophys J 1991;60:1438–44. [89] Söngen H, Bechstein R, Kühnle A. Quantitative atomic force microscopy. Journal of
[54] Butt HJ, Cappella B, Kappl M. Force measurements with the atomic force micro- Physics-Condensed Matter 2017;29(27):274001.
scope: technique, interpretation and applications. Surf Sci Rep 2005;59:1–152. [90] Watkins M, Shluger AL. Mechanism of contrast formation in atomic force microscopy
[55] Leiro J, Torhola M, Laajalehto K. The AFM method in studies of muscovite mica and in water. Phys Rev Lett 2010;105:196101.
galena surfaces. J Phys Chem Solid 2017;100:40–4. [91] Miyazawa K, Kobayashi N, Watkins M, Shluger AL, Amano K, Fukuma T. A relation-
[56] Beaussart A, Parkinson L, Mierczynska-Vasilev A, Beattie DA. Adsorption of modi- ship between three-dimensional surface hydration structures and force distribu-
fied dextrins on molybdenite: AFM imaging, contact angle, and flotation studies. J tion measured by atomic force microscopy. Nanoscale 2016;8:7334–42.
Colloid Interface Sci 2012;368:608–15. [92] Fukuma T, Onishi K, Kobayashi N, Matsuki A, Asakawa H. Atomic-raging in liquid by
[57] Xie L, Wang J, Shi C, Huang J, Zhang H, Liu Q, et al. Probing surface interactions of frequency modulation atomic force microscopy using small cantilevers with
electrochemically active galena mineral surface using atomic force microscopy. J megahertz-order resonance frequencies. Nanotechnology 2012;23:135706.
Phys Chem C 2016;120:22433–42. [93] Weber SA, Kilpatrick JI, Brosnan TM, Jarvis SP, Rodriguez BJ. High viscosity
[58] Lu Z, Liu Q, Xu Z, Zeng H. Probing anisotropic surface properties of molybdenite by environments: an unexpected route to obtain true atomic resolution with atomic
direct force measurements. Langmuir 2015;31:11409–18. force microscopy. Nanotechnology 2014;25(17):175701.
[59] Gupta V, Hampton MA, Nguyen AV, Miller JD. Crystal lattice imaging of the silica [94] Söngen H, Nalbach M, Adam H, Kühnle A. Three-dimensional atomic force micros-
and alumina faces of kaolinite using atomic force microscopy. J Colloid Interface copy mapping at the solid-liquid interface with fast and flexible data acquisition.
Sci 2010;352:75–80. Review of Scientific Instruments 2016;87:063704.
[60] Hampton MA, Plackowski C, Nguyen AV. Physical and chemical analysis of elemen- [95] Lohse D, Zhang XH. Surface nanobubbles and nanodroplets. Rev Mod Phys 2015;
tal sulfur formation during galena surface oxidation. Langmuir 2011;27:4190–201. 87:981–1035.
[61] Siretanu I, van den Ende D, Mugele F. Atomic structure and surface defects at [96] Zhang XH, Chan DYC, Wang DY, Maeda N. Stability of interfacial nanobubbles.
mineral-water interfaces probed by in situ atomic force microscopy. Nanoscale Langmuir 2013;29:1017–23.
2016;8(15):8220–7. [97] Schubert H. Nanobubbles, hydrophobic effect, heterocoagulation and hydrodynam-
[62] Gupta V, Miller JD. Surface force measurements at the basal planes of ordered kao- ics in flotation. Int J Miner Process 2005;78:11–21.
linite particles. J Colloid Interface Sci 2010;344:362–71. [98] Sobhy A, Tao D. Nanobubble column flotation of fine coal particles and associated
[63] Hayes RA, Ralston J. The collectorless flotation and separation of sulphide minerals fundamentals. Int J Miner Process 2013;124:109–16.
by Eh control. Int J Miner Process 1988;23(1–2):55–84. [99] Parker J, Claesson P, Attard P. Bubbles, cavities, and the long-ranged attraction be-
[64] Hayes RA, Price DM, Ralston J, Smith RW. Collectorless flotation of sulphide min- tween hydrophobic surfaces. J Phys Chem 1994;98:8468–80.
erals. Mineral Procesing and Extractive Metallurgy Review 1987;2(3):203–34. [100] Ishida N, Inoue T, Miyahara M, Higashitani K. Nano bubbles on a hydrophobic sur-
[65] Bruening F, Cohen A. Measuring surface properties and oxidation of coal macerals face in water observed by tapping-mode atomic force microscopy. Langmuir 2000;
using the atomic force microscope. Int J Coal Geol 2005;63:195–204. 16:6377–80.
[66] Morga R. Changes of semifusinite and fusinite surface roughness during heat treat- [101] Lou ST, Ouyang ZQ, Zhang Y, Li XJ, Hu J, Li MQ, et al. Nanobubbles on solid surface
ment determined by atomic force microscopy. Int J Coal Geol 2011;88:218–26. imaged by atomic force microscopy. Journal of Vacuum Science & Technology B:
[67] Pan JN, Zhu HT, Hou QL, Wang HC, Wang S. Macromolecular and pore structures of Microelectronics and Nanometer Structures Processing, Measurement, and Phe-
Chinese tectonically deformed coal studied by atomic force microscopy. Fuel 2015; nomena 2000;18:2573–5.
139:94–101. [102] Peng H, Hampton MA, Nguyen AV. Nanobubbles do not sit alone at the solid-liquid
[68] Wu D, Liu GJ, Sun RY, Chen SC. Influences of magmatic intrusion on the macromo- interface. Langmuir 2013;29:6123–30.
lecular and pore structures of coal: evidences from Raman spectroscopy and [103] Zhang XH, Maeda N, VSJ Craig. Physical properties of nanobubbles on hydrophobic
atomic force microscopy. Fuel 2014;119:191–201. surfaces in water and aqueous solutions. Langmuir 2006;22:5025–35.
[69] Drosthansen W. Effects of vicinal water on colloidal stability and sedimentation [104] Walczyk W, Schönherr H. Dimensions and the profile of surface nanobubbles: tip–
processes. J Colloid Interface Sci 1977;58:251–62. nanobubble interactions and nanobubble deformation in atomic force microscopy.
[70] Du H, Miller JD. A molecular dynamics simulation study of water structure and ad- Langmuir 2014;30:11955–65.
sorption states at talc surfaces. Int J Miner Process 2007;84:172–84. [105] Walczyk W, Schon PM, Schönherr H. The effect of PeakForce tapping mode AFM
[71] Gao J, Szoszkiewicz R, Landman U, Riedo E. Structured and viscous water in imaging on the apparent shape of surface nanobubbles. J Phys Condens Matter
subnanometer gaps. Physical Review B 2007;75:115415. 2013;25:184005.
[72] Fenter P, Lee SS. Hydration layer structure at solid-water interfaces. MRS Bulletin [106] Ko HC, Hsu WH, Yang CW, Fang CK, Lu YH, Hwang IS. High-resolution characteriza-
2014;39:1056–61. tion of preferential gas adsorption at the graphene-water interface. Langmuir
[73] Marutschke C, Walters D, Cleveland J, Hermes I, Bechstein R, Kühnle A. Three- 2016;32:11164–71.
dimensional hydration layer mapping on the (10.4) surface of calcite using ampli- [107] Walczyk W, Schönherr H. Characterization of the interaction between AFM tips and
tude modulation atomic force microscopy. Nanotechnology 2014;25:335703. surface nanobubbles. Langmuir 2014;30:7112–26.
[74] Miyazawa K, Watkins M, Shluger A, Fukuma T. Influence of ions on two- [108] Zhao B, Song Y, Wang S, Dai B, Zhang L, Dong Y, et al. Mechanical mapping of
dimensional and three-dimensional atomic force microscopy at fluorite-water in- nanobubbles by PeakForce atomic force microscopy. Soft Matter 2013;9:8837–43.
terfaces. Nanotechnology 2017;28:245701. [109] Schönherr H, Hain N, Walczyk W, Wesner D, Druzhinin SI. Surface nanobubbles
[75] Raviv U, Laurat P, Klein J. Fluidity of water confined to subnanometre films. Nature studied by atomic force microscopy techniques: Facts, fiction, and open questions.
2001;413:51–4. Japanese Journal of Applied Physics 2016;55:08NA01.
[76] Raviv U, Perkin S, Laurat P, Klein J. Fluidity of water confined down to [110] Miller JD, Hu YH, Veeramasuneni S, Lu YQ. In-situ detection of butane gas at a hy-
subnanometer films. Langmuir 2004;20:5322–32. drophobic silicon surface. Colloids Surf A Physicochem Eng Asp 1999;154:137–47.
[77] Björneholm O, Hansen MH, Hodgson A, Liu LM, Limmer DT, Michaelides A, et al. [111] Chan CU, Ohl CD. Total-internal-reflection-fluorescence microscopy for the study of
Water at interfaces. Chem Rev 2016;116(13):7698. nanobubble dynamics. Phys Rev Lett 2012;109:174501.
[78] Söngen H, Marutschke C, Spijker P, Holmgren E, Hermes IM, Bechstein R, et al. [112] Li M, Tonggu L, Zhan X, Mega TL, Wang L. Cryo-EM visualization of nanobubbles in
Chemical indentification at the solid-liquid interface. Langmuir 2017;33(1):125–9. aqueous solutions. Langmuir 2016;32:11111–5.
[79] Fukuma T, Reischl B, Kobayashi N, Spijker P, Canova FF, Miyazawa K, et al. Mecha- [113] Switkes M, Ruberti J. Rapid cryofixation/freeze fracture for the study of
nism of atomic force microscopy imaging of three-dimensional hydration struc- nanobubbles at solid–liquid interfaces. Appl Phys Lett 2004;84:4759–61.
tures at a solid-liquid interface. Physical Review B 2015;92:155412. [114] Zhang XH. Quartz crystal microbalance study of the interfacial nanobubbles. Phys
[80] Mezger M, Reichert H, Schöder S, Okasinski J, Schröder H, Dosch H, et al. High- Chem Chem Phys 2008;10:6842–8.
resolution in situ x-ray study of the hydrophobic gap at the water-octadecyl- [115] Hain N, Wesner D, Druzhinin SI, Schönherr H. Surface nanobubbles studied by
trichlorosilane interface. Proc Natl Acad Sci 2006;103:18401–4. time-resolved fluorescence microscopy methods combined with AFM: the impact
[81] Doshi DA, Watkins EB, Israelachvili JN, Majewski J. Reduced water density at hydro- of surface treatment on nanobubble nucleation. Langmuir 2016;32:11155–63.
phobic surfaces: effect of dissolved gases. Proc Natl Acad Sci 2005;102:9458–62. [116] Pan G, He G, Zhang M, Zhou Q, Tyliszczak T, Tai R, et al. Nanobubbles at hydrophilic
[82] Jensen TR, Jensen MØ, Reitzel N, Balashev K, Peters GH, Kjaer K, et al. Water in con- particle-water interfaces. Langmuir 2016;32:11133–7.
tact with extended hydrophobic surfaces: direct evidence of weak dewetting. Phys [117] Palmer LA, Cookson D, Lamb RN. The relationship between nanobubbles and the
Rev Lett 2003;90:086101. hydrophobic force. Langmuir 2011;27:144–7.
[83] Fukuma T, Kobayashi K, Matsushige K, Yamada H. True atomic resolution in liquid by [118] Alheshibri M, Qian J, Jehannin M, Craig VS. A history of nanobubbles. Langmuir
frequency-modulation atomic force microscopy. Appl Phys Lett 2005;87:034101. 2016;32:11086–100.
[84] Zachariah Z, Espinosa-Marzal RM, Spencer ND, Heuberger MP. Stepwise collapse of [119] Zhang XH, Zhang XD, Lou ST, Zhang ZX, Sun JL, Degassing Hu J. Temperature effects
highly overlapping electrical double layers. Phys Chem Chem Phys 2016;18: on the formation of nanobubbles at the mica/water interface. Langmuir 2004;20
24417–27. (9):3813–5.
[85] Fukuma T, Ueda Y, Yoshioka S, Asakawa H. Atomic-scale distribution of water mol- [120] Zhang XH, Zhang X, Sun J, Zhang Z, Li G, Fang H, et al. Detection of novel gaseous
ecules at the mica-water interface visualized by three-dimensional scanning force states at the highly oriented pyrolytic graphite-water interface. Langmuir 2007;
microscopy. Phys Rev Lett 2010;104:016101. 23:1778–83.
[86] Imada H, Kimura K, Onishi H. Water and 2-propanol structured on calcite (104) [121] van Limbeek MA, Seddon JR. Surface nanobubbles as a function of gas type.
probed by frequency-modulation atomic force microscopy. Langmuir 2013;29: Langmuir 2011;27:8694–9.
10744–51. [122] Zhang X, Uddin MH, Yang H, Toikka G, Ducker W, Maeda N. Effects of surfactants on
[87] Sader JE, Jarvis SP. Accurate formulas for interaction force and energy in frequency the formation and the stability of interfacial nanobubbles. Langmuir 2012;28:
modulation force spectroscopy. Appl Phys Lett 2004;84(10):1801–3. 10471–7.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx 19

[123] Xu C, Peng S, Qiao GG, Gutowski V, Lohse D, Zhang X. Nanobubble formation on a [155] Christensen H, Claesson P. Cavitation and the interaction between macroscopic sur-
warmer substrate. Soft Matter 2014;10:7857–64. faces. Science 1988;239:390–2.
[124] Seiedi O, Rahbar M, Nabipour M, Emadi MA, Ghatee MH, Ayatollahi S. Atomic force [156] Xing YW, Xu XH, Gui XH, Cao YJ, Xu MD. Effect of kaolinite and montmorillonite on
microscopy (AFM) investigation on the surfactant wettability alteration mecha- fine coal flotation. Fuel 2017;195:284–9.
nism of aged mica mineral surfaces. Energy Fuel 2010;25:183–8. [157] Shrimali K, Jin J, Hassas BV, Wang X, Miller JD. The surface state of hematite and its
[125] Fielden ML, Claesson PM, Verrall RE. Investigating the adsorption of the gemini wetting characteristics. J Colloid Interface Sci 2016;477:16–24.
surfactant “12-2-12” onto mica using atomic force microscopy and surface force [158] Borkovec M, Papastavrou G. Interactions between solid surfaces with adsorbed
apparatus measurements. Langmuir 1999;15:3924–34. polyelectrolytes of opposite charge. Curr Opin Colloid Interface Sci 2008;13:
[126] Dong J, Mao G. Direct study of C12E5 aggregation on mica by atomic force micros- 429–37.
copy imaging and force measurements. Langmuir 2000;16:6641–7. [159] Borkovec M, Szilagyi I, Popa I, Finessi M, Sinha P, Maroni P, et al. Investigating forces
[127] Ferrari M, Ravera F, Viviani M, Liggieri L. Characterization of surfactant aggregates between charged particles in the presence of oppositely charged polyelectrolytes
at solid–liquid surfaces by atomic force microscopy. Colloids Surf A Physicochem with the multi-particle colloidal probe technique. Adv Colloid Interface Sci 2012;
Eng Asp 2004;249:63–7. 179:85–98.
[128] Chennakesavulu K, Raju GB, Prabhakar S, Nair CM, Murthy K. Adsorption of oleate [160] Abraham T, Christendat D, Xu Z, Masliyah J, Gohy J-F, Jérôme R. Role of polyelectro-
on fluorite surface as revealed by atomic force microscopy. Int J Miner Process lyte charge density in tuning colloidal forces. AIChE Journal 2004;50:2613–26.
2009;90:101–4. [161] Zhou Y, Gan Y, Wanless EJ, Jameson GJ, Franks GV. Interaction forces between silica
[129] Paiva P, Monte M, Simao R, Gaspar J. In situ AFM study of potassium oleate adsorp- surfaces in aqueous solutions of cationic polymeric flocculants: effect of polymer
tion and calcium precipitate formation on an apatite surface. Miner Eng 2011;24: charge. Langmuir 2008;24:10920–8.
387–95. [162] Ofori P, Nguyen AV, Firth B, McNally C, Hampton MA. The role of surface interaction
[130] Manne S. Molecular organization of surfactants at solid-liquid interfaces. Science forces and mixing in enhanced dewatering of coal preparation tailings. Fuel 2012;
1995;270(5241):1480–2. 97:262–8.
[131] Warr GG. Surfactant adsorbed layer structure at solid/solution interfaces: impact [163] Ruiz-Cabello FJM, Maroni P, Borkovec M. Direct measurements of forces between
and implications of AFM imaging studies. Curr Opin Colloid Interface Sci 2000;5 different charged colloidal particles and their prediction by the theory of Derjaguin,
(1–2):88–94. Landau, Verwey, and Overbeek (DLVO). J Chem Phys 2013;138:234705.
[132] Paruchuri VK, Nguyen AV, Miller JD. Zeta-potentials of self-assembled surface mi- [164] Ivanova NO, Xu Z, Liu Q, Masliyah JH. Surface forces in unconventional oil process-
celles of ionic surfactants adsorbed at hydrophobic graphite surfaces. Colloids ing. Curr Opin Colloid Interface Sci 2017;27:63–73.
and Surfaces a-Physicochemical and Engineering Aspects 2004;250(1–3):519–26. [165] Masliyah J, Czarnecki J, Xu Z. Handbook on theory and practice of bitumen recovery
[133] Paruchuri VK, Nalaskowski J, Shah DO, Miller JD. The effect of cosurfactants on so- from Athabasca oil sands. Theoretical Basis 2011;1.
dium dodecyl sulfate micellar structures at a graphite surface. Colloids and Surfaces [166] Liu J, Xu Z, Masliyah J. Interaction forces in bitumen extraction from oil sands. J Col-
a-Physicochemical and Engineering Aspects 2006;272(3):157–63. loid Interface Sci 2005;287:507–20.
[134] Jaschke M, Butt HJ, Graub HE, Manne S. Surfactant Aggregates at a metal surface. [167] Long J, Drelich J, Xu Z, Masliyah JH. Effect of operating temperature on water-based
Langmuir 1997;13(13):1381–4. oil sands processing. Can J Chem Eng 2007;85:726–38.
[135] Xie L, Wang J, Yuan D, Shi C, Cui X, Zhang H, et al. Interaction mechanisms between [168] Zhang Y, Ding M, Liu J, Jia W, Ren S. Studies on bitumen–silica interaction in surfac-
air bubble and molybdenite surface: impact of solution salinity and polymer ad- tants and divalent cations solutions by atomic force microscopy. Colloids Surf A
sorption. Langmuir 2017;33:2353–61. Physicochem Eng Asp 2015;482:241–7.
[136] Mierczynska-Vasilev A, Beattie DA. Adsorption of tailored carboxymethyl cellulose [169] Liu J, Xu Z, Masliyah J. Role of fine clays in bitumen extraction from oil sands. AIChE
polymers on talc and chalcopyrite: correlation between coverage, wettability, and Journal 2004;50:1917–27.
flotation. Miner Eng 2010;23:985–93. [170] Diao M, Taran E, Mahler S, Nguyen AV. A concise review of nanoscopic aspects of
[137] Rief M, Oesterhelt F, Heymann B, Gaub HE. Single molecule force spectroscopy on bioleaching bacteria-mineral interactions. Adv Colloid Interface Sci 2014;212:
polysaccharides by atomic force microscopy. Science 1997;275:1295–7. 45–63.
[138] Neuman KC, Nagy A. Single-molecule force spectroscopy: optical tweezers, mag- [171] Meyer RL, Zhou X, Tang L, Arpanaei A, Kingshott P, Besenbacher F. Immobilisation
netic tweezers and atomic force microscopy. Nat Methods 2008;5:491–505. of living bacteria for AFM imaging under physiological conditions. Ultramicroscopy
[139] Liu C, Shi W, Cui S, Wang Z, Zhang X. Force spectroscopy of polymers: beyond sin- 2010;110:1349–57.
gle chain mechanics. Current Opinion in Solid State and Materials Science 2005;9: [172] Bowen WR, Hilal N, Lovitt RW, Wright CJ. Direct measurement of the force of adhe-
140–8. sion of a single biological cell using an atomic force microscope. Colloids Surf A
[140] Zhang X, Liu C, Wang Z. Force spectroscopy of polymers: studying on intramolecu- Physicochem Eng Asp 1998;136:231–4.
lar and intermolecular interactions in single molecular level. Polymer 2008;49: [173] Lower SK, Tadanier CJ, Hochella MF. Measuring interfacial and adhesion forces be-
3353–61. tween bacteria and mineral surfaces with biological force microscopy. Geochim
[141] Sun W, Long J, Xu Z, Masliyah JH. Study of Al(OH)3-polyacrylamide-induced Cosmochim Acta 2000;64:3133–9.
pelleting flocculation by single molecule force spectroscopy. Langmuir 2008;24: [174] Diao M, Nguyen TA, Taran E, Mahler S, Nguyen AV. Differences in adhesion of
14015–21. A. thiooxidans and A. ferrooxidans on chalcopyrite as revealed by atomic force mi-
[142] Long J, Xu Z, Masliyah JH. Adhesion of single polyelectrolyte molecules on silica, croscopy with bacterial probes. Miner Eng 2014;61:9–15.
mica, and bitumen surfaces. Langmuir 2006;22:1652–9. [175] Chan DY, Klaseboer E, Manica R. Theory of non-equilibrium force measurements in-
[143] Pensini E, Yip CM, O'Carroll D, Carboxymethyl Sleep BE. Cellulose binding to min- volving deformable drops and bubbles. Adv Colloid Interface Sci 2011;165:70–90.
eral substrates: characterization by atomic force microscopy–based force spectros- [176] Ducker WA, Xu Z, Israelachvili JN. Measurements of hydrophobic and DLVO forces
copy and quartz-crystal microbalance with dissipation monitoring. J Colloid in bubble-surface interactions in aqueous solutions. Langmuir 1994;10:3279–89.
Interface Sci 2013;402:58–67. [177] Butt H-J. A technique for measuring the force between a colloidal particle in water
[144] Fa K, Nguyen AV, Miller JD. Interaction of calcium dioleate collector colloids with and a bubble. J Colloid Interface Sci 1994;166:109–17.
calcite and fluorite surfaces as revealed by AFM force measurements and molecular [178] Tabor RF, Grieser F, Dagastine RR, Chan DY. Measurement and analysis of forces in
dynamics simulation. Int J Miner Process 2006;81:166–77. bubble and droplet systems using AFM. J Colloid Interface Sci 2012;371:1–14.
[145] Fa K, Jiang T, Nalaskowski J, Miller JD. Interaction forces between a calcium dioleate [179] Johnson DJ, Miles NJ, Hilal N. Quantification of particle-bubble interactions using
sphere and calcite/fluorite surfaces and their significance in flotation. Langmuir atomic force microscopy: a review. Adv Colloid Interface Sci 2006;127:67–81.
2003;19:10523–30. [180] Nguyen AV, Nalaskowski J, Miller JD. A study of bubble-particle interaction using
[146] Free ML, Miller JD. Kinetics of 18-carbon carboxylate adsorption at the fluorite sur- atomic force microscopy. Miner Eng 2003;16:1173–81.
face. Langmuir 1997;13:4377–82. [181] Assemi S, Nguyen AV, Miller JD. Direct measurement of particle–bubble interaction
[147] Xing Y, Li C, Gui X, Cao Y. Interaction forces between paraffin/stearic acid and fresh/ forces using atomic force microscopy. Int J Miner Process 2008;89(1):65–70.
oxidized coal particles measured by atomic force microscopy. Energy Fuel 2017;31: [182] Nguyen AV, Evans GM, Nalaskowski J, Miller JD. Hydrodynamic interaction
3305–12. between an air bubble and a particle: atomic force microscopy measurements.
[148] Xing Y, Gui XH, Cao YJ. Effect of calcium ion on coal flotation in the presence of ka- Experimental Thermal and Fluid Science 2004;28:387–94.
olinite clay. Energy Fuel 2016;30(2):1517–23. [183] Ishida N. Direct measurement of hydrophobic particle-bubble interactions in
[149] Gui XH, Xing YW, Rong GQ, Cao YJ, Liu JT. Interaction forces between coal and ka- aqueous solutions by atomic force microscopy: effect of particle hydrophobicity.
olinite particles measured by atomic force microscopy. Powder Technol 2016;301: Colloids Surf A Physicochem Eng Asp 2007;300:293–9.
349–55. [184] Fielden ML, Hayes RA, Ralston J. Surface and capillary forces affecting air bubble−
[150] Forbes E. Shear, selective and temperature responsive flocculation: a comparison of particle interactions in aqueous electrolyte. Langmuir 1996;12:3721–7.
fine particle flotation techniques. Int J Miner Process 2011;99(1–4):1–10. [185] Englert A, Krasowska M, Fornasiero D, Ralston J, Rubio J. Interaction force between
[151] Rabinovich YI, Yoon RH. Use of atomic-force microscope for the measurements of an air bubble and a hydrophilic spherical particle in water, measured by the colloid
hydrophobic forces between silanated silica plate and glass sphere. Langmuir probe technique. Int J Miner Process 2009;92:121–7.
1994;10:1903–9. [186] Taran E, Hampton MA, Nguyen AV, Attard P. Anomalous time effect on particle-
[152] Tyrrell JWG, Attard P. Atomic force microscope images of nanobubbles on a hydro- bubble interactions studied by atomic force microscopy. Langmuir 2009;25:
phobic surface and corresponding force-separation data. Langmuir 2002;18:160–7. 2797–803.
[153] Hampton MA, Nguyen AV. Systematically altering the hydrophobic nanobubble [187] Scheludko A. Attachment of particles to a liquid surface (capillary theory of flota-
bridging capillary force from attractive to repulsive. J Colloid Interface Sci 2009; tion). Journal of the Chemical Society Faraday Transactions 1976;72(4):2815–28.
333:800–6. [188] James DF. The meniscus on the outside of a small circular cylinder. J Fluid Mech
[154] Hampton MA, Donose BC, Nguyen AV. Effect of alcohol-water exchange and surface 2006;63(4):657–64.
scanning on nanobubbles and the attraction between hydrophobic surfaces. J Col- [189] O'Brien SBG. The meniscus near a small sphere and its relationship to line pinning
loid Interface Sci 2008;325:267–74. of contact lines. Journal of Colloid & Interface Science 1996;183(1):51–6.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004
20 Y. Xing et al. / Advances in Colloid and Interface Science xxx (2018) xxx–xxx

[190] Chateau X, Pitois O. Quasistatic detachment of a sphere from a liquid interface. [207] Schönherr H. Forces and thin water film drainage in deformable asymmetric nano-
Journal of Colloid & Interface Science 2003;259(2):346–53. scale contacts. ACS Nano 2015;9:12–5.
[191] Butt H, Gao N, Papadopoulos P, Steffen W, Kappl M, Berger R. Energy dissipation of [208] Yin X, Gupta V, Du H, Wang X, Miller JD. Surface charge and wetting characteristics
moving drops on superhydrophobic and superoleophobic surfaces. Langmuir 2017; of layered silicate minerals. Adv Colloid Interface Sci 2012;179:43–50.
33(1). [209] Hartley PG, Larson A, Scales PJ. Electrokinetic amd direct force measurements
[192] Mcnamee CE, Kappl M, Butt HJ, Higashitani K, Graf K. Interfacial forces between a between silica and mica surfaces in dilute electrolyte solutions. Langmuir 1997;
silica particle and phosphatidylcholine monolayers at the air-water interface. 13(8):2207–14.
Langmuir 2010;26(18):14574–81. [210] Assemi S, Nalaskowski J, Miller JD, Johnson WP. Isoelectric point of fluorite by di-
[193] Mcnamee CE, Kappl M, Butt H, Ally J, Shigenobu H, Iwafuji Y, et al. Forces between a rect force measurements using atomic force microscopy. Langmuir 2006;22(4):
monolayer at an air/water interface and a particle in solution: influence of the 1403–5.
sign of the surface charges and the subphase salt concentration. Soft Matter [211] Butt HJ. Measuring local surface charge densities in electrolyte solutions with a
2011;7(21):10182–92. scanning force microscope. Biophys J 1992;63(2):578–82.
[194] Anachkov SE, Lesov I, Zanini M, Kralchevsky PA, Denkov ND, Isa L. Particle detachment [212] And CR, Radmacher M. Mapping local electrostatic forces with the atomic force mi-
from fluid interfaces: theory vs. experiments. Soft Matter 2016;12(36):7632–43. croscope. Langmuir 1997;13(10):2825–32.
[195] Preuss M, Butt HJ. Direct measurement of particle-bubble interactions in aqueous [213] Zhao H, Bhattacharjee S, Chow R, Wallace D, Masliyah JH, Xu Z. Probing surface
electrolyte: dependence on surfactant. Langmuir 1998;14:3164–74. charge potentials of clay basal planes and edges by direct force measurements.
[196] Preuss M, Butt HJ. Direct measurement of forces between particles and bubbles. Int Langmuir 2008;24:12899–910.
J Miner Process 1999;56:99–115. [214] Drelich J, Long J, Yeung A. Determining surface potential of the bitumen-water in-
[197] Nguyen AV, Nalaskowski J, Miller JD. The dynamic nature of contact angles as mea- terface at nanoscale resolution using atomic force microscopy. Can J Chem Eng
sured by atomic force microscopy. J Colloid Interface Sci 2003;262(1):303–6. 2007;85:625–34.
[198] Chan DY, Manor O, Connor JN, Horn RG. Soft matter: from shapes to forces on the [215] Yan L, Englert AH, Masliyah JH, Xu Z. Determination of anisotropic surface charac-
nanoscale. Soft Matter 2008;4:471–4. teristics of different phyllosilicates by direct force measurements. Langmuir 2011;
[199] Manica R, Hendrix MHW, Gupta R, Klaseboer E, Ohl CD, DYC Chan. Effects of hydrody- 27:12996–3007.
namic film boundary conditions on bubble-wall impact. Soft Matter 2013;9:9755–8. [216] Gao Z, Hu Y, Sun W, Drelich JW. Surface-charge anisotropy of scheelite crystals.
[200] Shi C, Chan DYC, Liu QX, Zeng HB. Probing the hydrophobic interaction between air Langmuir 2016;32:6282–8.
bubbles and partially hydrophobic surfaces using atomic force microscopy. J Phys [217] Xie L, Wang J, Shi C, Cui X, Huang J, Zhang H, et al. Mapping the nanoscale hetero-
Chem C 2014;118:25000–8. geneity of surface hydrophobicity on the sphalerite mineral. J Phys Chem C 2017;
[201] Xie L, Shi C, Wang J, Huang J, Lu Q, Liu Q, et al. Probing the interaction between air 121:5620–8.
bubble and sphalerite mineral surface using atomic force microscope. Langmuir [218] Fritzsche J, Peuker UA. Particle adhesion on highly rough hydrophobic surfaces: the
2015;31:2438–46. distribution of interaction mechanisms. Colloids Surf A Physicochem Eng Asp 2014;
[202] Clark SC, Walz JY, Ducker WA. Atomic force microscopy colloid-probe measure- 459:166–71.
ments with explicit measurement of particle-solid separation. Langmuir 2004; [219] Rudolph M, Peuker UA. Hydrophobicity of minerals determined by atomic force
20(18):7616–22. microscopy - a tool for flotation research. Chemie Ingenieur Technik 2014;86:
[203] McKee CT, Clark SC, Walz JY, Ducker WA. Relationship between scattered intensity 865–73.
and separation for particles in an evanescent field. Langmuir 2005;21:5783–9. [220] Rudolph M, Peuker UA. Mapping hydrophobicity combining AFM and Raman spec-
[204] McKee CT, Mosse WKJ, Ducker WA. Measurement of the absolute separation for troscopy. Minerals Engineering 2014;66–68:181–90.
atomic force microscopy measurements in the presence of adsorbed polymer. Re- [221] Dazzi A, Prater CB. AFM-IR: technology and applications in nanoscale infrared spec-
view of scientific instruments 2006;77:053706. troscopy and chemical imaging. Chem Rev 2016;117:5146–73.
[205] McKee CT. Investigation of non-DLVO forces using an evanescent wave atomic [222] Fu WY, Zhang W. Hybrid AFM for nanoscale physicochemical characterization: re-
force microscope: Virginia Polytechnic Institute and State University; 2006. cent development and emerging applications. Small 2017;13(11):1603525.
[206] Shi C, Cui X, Xie L, Liu QX, Chan DYC, Israelachvili JN, et al. Measuring forces and [223] Eifert A, Kranz C. Hyphenating atomic force microscopy. Anal Chem 2014;86(11):
spatiotemporal evolution of thin water films between an air bubble and solid sur- 5190–200.
faces of different hydrophobicity. ACS Nano 2015;9:95–104.

Please cite this article as: Xing Y, et al, The application of atomic force microscopy in mineral flotation, Adv Colloid Interface Sci (2018), https://doi.
org/10.1016/j.cis.2018.01.004

You might also like