Sunteți pe pagina 1din 27

Dispersion Behavior of -Cyclodextrin Nanobubbles

by

Treiana Mari S. Collado

A Thesis Proposal Submitted to the


School of Chemical, Biological, and Materials Engineering and Sciences
in Partial Fulfillment of the Requirements for the Degree

Bachelor of Science in Chemical Engineering


Mapúa University
July 2020
TABLE OF CONTENTS

TITLE PAGE i

APPROVAL PAGE ii

ACKNOWLEDGEMENT iii

TABLE OF CONTENTS iv

LIST OF TABLES vi

LIST OF FIGURES vii

ABSTRACT viii

Chapter 1: INTRODUCTION 1

Chapter 2: REVIEW OF LITERATURE 5

1. Heading 1 5
2. Heading 2 9
3. Heading 3 15

TITLE OF A MAJOR ASPECT OF STUDY 21

1. Methodology 21

REFERENCES 30
LIST OF TABLES

Chapter 2
TABLE 1:
TABLE 2:
LIST OF FIGURES

Chapter 2
FIGURE 1: TITLE 16
FIGURE 2: TITLE … 23
FIGURE 3: TITLE 24
FIGURE 4: TITLE … 27

Chapter 3
FIGURE 1: TITLE 16
FIGURE 2: TITLE … 23
FIGURE 3: TITLE 24
FIGURE 4: TITLE … 27
Chapter 1

INTRODUCTION

In 1891, “cellulosines” or what is recognized today as cyclodextrins are first coined

by A. Villiers, a French pharmacist and a chemist [1]. A few years after, Franz Schardinger,

an Austrian scientist, discovered the  three naturally forms of the compound or the

“Schardinger sugars”, pertaining to cyclodextrins -α, -β, and -γ, granting him the title

“Founding Founder” of cyclodextrins [2]. A family of cyclic oligosaccharides, comprising of

a macrocyclic ring of glucose subunits joined by α-1,4 glycosidic bonds, cyclodextrins are

derivatives of starch through enzymatic conversion [3].

β-Cyclodextrin is a cyclic derivative of starch prepared from partially hydrolyzed

starch (maltodextrin) by an enzymatic process. Cyclodextrins are a group of naturally

occurring cyclic oligosaccharides with six, seven, or eight glucose residues linked by α(1 →

4) glycoside bonds in a cylindrical structure that are denominated α-, β-, and γ-cyclodextrins,

respectively[3].

Engineered nanomaterials (NMs) are widely used in many consumer and healthcare

products, as well as novel nanomedicines.1 To enable a quick and reliable NMs risk-benefit

assessment, there is an urgent need for robust, standardised and reliable methods for their

characterisation and toxicity screening. For this purpose, the understanding of the behaviour

and fate of these NMs when in contact with biological systems is important. Physical and

chemical properties such as size, surface chemistry and surface charge have been identified

as essential parameters to determine, because they affect the NMs’ mode of action in a given
environment (water, buffer, biological fluid, etc.) through different surface molecular

interactions [4]

Nanobubbles or ultrafine bubbles are defined as cavities of gases with diameter

<200nm in aqueous solutions [5]. Industrial application of nano- bubbles has exponentially

increased over the past two de- cades due to their reactivity and stability, compared with

macro- and microbubbles. Due to the size, they have high specific surface areas and high

stagnation times [6] which increases mass transport efficiencies, physical absorptions, and

chemical reactions at the gas–liquid inter- faces. Moreover, these bubbles have long

residence time in solutions and electrically charged surfaces [7]. Due to the above,

nanobubbles have many industrial applications such as manufacturing of functional

materials, soil and sediment decontamination, pharmaceutical delivery, and disinfection of

food products [8]. After generation, nanobubbles are found to exist in aqueous solutions for

several weeks. It is reported that bubbles of radii 150–200nm were in a solution for 2 weeks

[9]. The electrically charged liquid–gas interface of nanobubbles create repulsive forces that

prevent bubble coalescence and, hence, high bubble densities creating highly dissolved gas

concentrations in water creating smaller con- centration gradients between the interface and

the bulk liquid [10]. Moreover, the stability of bubbles increased by low rising velocity,

which is negligible due to Brownian motion and low buoyancy forces [5]. Other than these

reasons, nanobubbles are considered to be stable by a mutual shielding against the diffusive

outflow of gases, which can be achieved if bubbles are sufficiently close together or gathered

into micrometer-sized clusters [11]. As of now, presence of stable nanobubbles has been

experimentally confirmed, yet a clear theoretical basis has not been established to explain

their long-term stability. Hence for effective and functional use of these bubbles, knowing
their properties and behavior is quite important. Yet, nano- bubble behavior is considered to

be complex. Therefore, a further study is required for proper understanding of the stability of

nanobubbles, impacted by generation techniques, coalescence, free radical generation, and

influencing factors of stability of nanobubbles, such as temperature, pressure, pH, salt

concentration, ion strength, presence of organic matters, the origin of the nanobubbles, and

so on..

A new method has been described for the separation of NMs according to their

hydrophobicity based on a set of collectors composed of fluorinated hydrophobic surfaces

whose surface energy components can be modified and fine- tuned by the layer-by-layer

(LbL) deposition of polyelectrolytes [4]. A proof-of-concept study of the method was carried

out with hydrophobic and hydrophilic polystyrene nanoparticles. The experimental results

were qualitatively supported by the eXtended Derjaguin Landau Verwey Overbeek

(XDLVO) theory, enabling the assessment of the different forces in play. The main objective

of this study was to investigate the long-term stability of nanobubbles based on factors that

influence the generation of nanobubbles. Therefore, nanobubbles were generated under the

following four different conditions and tested for their size distributions and zeta potential

values. Then, the results are to be analyzed.

The impacts of the infilled gas types were investigated. The test results showed that

the size and zeta potential values of nano oxygen, nitrogen, air, and ozone bubbles were the

function of the properties of the gas type, specifically the gas solubility. The nitrogen gas

with the least solubility had the smallest bubble diameter, while ozone with the highest gas

solubility produced the largest diameter bubbles. The negative zeta potential value of

nanobubbles is due to number of OH- ions on the bubble surface. Since all the parameters are
identical except the gas type of nanobubbles, it can be concluded that the zeta potential is a

function of the gas diffusion rates and gas solubility and would contribute to the generation

of OH- ions on the bubble surface. Based on test results, ozone had highest magnitude

negative zeta potential value followed by oxygen, air, and nitrogen [12].

As of now, presence of stable nanobubbles has been experimentally confirmed, yet a

clear theoretical basis has not been established to explain their long-term stability. Hence for

effective and functional use of these bubbles, knowing their properties and behavior is quite

important. Yet, nanobubble behavior is considered to be complex. Therefore, a further study

is required for proper understanding of the stability of nanobubbles, impacted by generation

techniques, coalescence, free radical generation, and influencing factors of stability of

nanobubbles, such as temperature, pressure, pH, salt concentration, ion strength, presence of

organic matters, the origin of the nanobubbles, and so on. The following section describes the

exponential growth of the application of nanobubbles in many fields of science, technology,

and industry. [4]

Many reports in the literature discuss about the difficulty of measuring the zeta

potential in a dispersion of bubbles because of the interference of the gravitational field and

of the lack of stability of the bubbles during the measurements [13]. In the case of MNB, the

long residence time in the liquid due to the low rising velocity of the bubble is an advantage

in the zeta potential measurement [14].

Similarly, zeta potential of NB produced with different kind of gases are measured.

The aim was to understand the influence of the type of gas in the MNB surface charge.

This research aims to investigate and analyze the impacts of hydrophobicity, overall

stability, pH and zeta potential values, and as well as other relevant properties on the
dispersion mecheanism of -cyclodextrin nanobubbles. Other possibilities such as deviations

from bulk formation of nanobubbles are ruled out despite of the similarities in methods of

analyzation.
Chapter 2

REVIEW OF RELATED LITERATURE

1. Contact Angle Measurement and Quantification of Nanoparticle Surface Hydrophobicity

1.1. Contact Angle and Hydrophobicity

The action of flotation is described by an attachment of a particle to air bubbles. Upon

the suspension of minerals in water, the air bubble displaces water from the surface of the

particle, making an effective contact that occurs when the mineral surface is or has been

made hydrophobic to considerable degree. The hydrophobicity of a particle can be

defined as the measure which the air bubble displaces water for the solid surface and can

be studied upon the through analyzing of the thermodynamic aspects of the bubble-

particle interface. [15]

1.2 Contact Angle Measurement and Hydrophobicity Quantification for Nanoparticle

Hydrophobicity is an important parameter for the risk assessment of chemicals, but

standardised quantitative methods for the determination of hydrophobicity cannot be

applied to nanomaterials. Here, a method for the direct quantification of the surface

energy and hydrophobicity of nanomaterial is defined. The quantification is obtained by

comparing the nanomaterial binding affinity to two or more engineered collectors, i.e.

surfaces with tuned hydrophobicity. In order to validate the concept, the method is

applied to a set of nanoparticles with varying degrees of hydrophobicity. The technique

described represents an alternative to the use of other methods such as hydrophobic


interaction chromatography or water–octanol partition, which provide only qualitative

values of hydrophobicity. [4]

1.3 Implications of Hydrophobicity on the Formation of Nanobubbles

The formation of nanobubbles at solid-liquid interfaces has been studied using the

atomic force microscopy (AFM) imaging technique. Nanobubble formation strongly

depends on both the hydrophobicity of the solid surface and the polarity of the liquid

subphase [16] . First, a number of homogenous surfaces are used as solid substrates to

understand and analyze the formation and distribution of nanobubbles on homogenous

surfaces with differing degrees of hydrophobicity. While nanobubbles do not form on flat

hydrophilic surfaces immersed in water, they appear spontaneously at the interface of

water against smooth, hydrophobic surfaces. From the experimental evidence, it is

concluded that the features observed in the AFM images are deformable, air-filled

bubbles [16]. In addition to the hydrophobicity of the solid surface, differences in

solubility of air between two miscible fluids can also lead to nucleation and growth of

nanobubbles. We observe that nanobubbles appear at the interface of water against

hydrophilic silicon oxide surfaces after in-situ mixing of ethanol and water in the fluid-

cell.

While for most part, the shapes of the nanobubbles are well approximated by

spherical caps with width much larger than the height; in the vicinity of the three phase

contact line, the shapes deviate significantly from the spherical cap profile to merge with

the solid surface at a slope of < 0.05. This deviation in the interfacial profile from the

spherical cap shape is due to long-range van der Waals forces, which are relevant at a
spatial scale of few nanometers from the solid surface. We quantify the morphological

distribution of nanobubbles by evaluating several important bubble parameters including

surface coverage and radii of curvature. In conjunction, with an analytical model

available in the literature, we use this information to estimate that the present nanobubble

morphology may give rise to slip lengths that are approximately 1-2 pm in pressure

driven flows for water flowing over a typical hydrophobic surface. The consistency of the

calculated slip length with experimental values reported recently in the literature,

suggests that the apparent fluid slip observed experimentally at hydrophobic surfaces

may indeed arise from the presence of nanobubbles. Further, trends are established

between different morphological parameters and surface hydrophobicity. Bubbles are

found to get bigger, wider and less-frequent in number with increasing surface

hydrophobicity. The pressure inside the bubble is found to decrease with an increase in

the surface hydrophobicity [16].

Contact angle of solid-liquid surfaces is crucial for nanoscale gas state formation

and its stability. The amount of form nanoscale gas state was increased with the increase

of surface area of hydrophobic surface area. For hydrophobic surfaces residual gas may

directly nucleate and accumulate on the surfaces the form surface none of scale gas state.

On the other hand, for hydrophilic surfaces, nanoscale gas state and also be form in the

services but that can quickly be destabilized to form microbubbles and eventually escape

from the surfaces. the stability of nanoscale gas state can vary greatly depending on the

hydrophobicity of the solid surfaces. These observations may facilitate further

development of nanobubble theory and the application of nanobubble technologies in

related areas. [17]


2. Stability of Nanobubbles in Water

After generation, nanobubbles are found to exist in aqueous solutions for several

weeks. It is reported that bubbles of radii 150–200 nm were in a solution for 2 weeks. The

electrically charged liquid–gas interface of nanobubbles create repulsive forces that prevent

bubble coalescence and, hence, high bubble densities creating highly dissolved gas

concentrations in water creating smaller concentration gradients between the interface and

the bulk liquid . Moreover, the stability of bubbles increased by low rising velocity, which is

negligible due to Brownian motion and low buoyancy forces [5]Other than these reasons,

nanobubbles are considered to be stable by a mutual shielding against the diffusive outflow

of gases, which can be achieved if bubbles are sufficiently close together or gathered into

micrometer-sized clusters [11].

2.1 Variations in Solution pH and Its Relation to Zeta Potentials and Bubble Size

Figure 1a shows the variation in zeta potential values with solution pH values,

where with increased solution pH, negativity of zeta potential increases. Figure 1b

shows the change in bubble size with solution pH values, where with higher solution

pH values, smaller bubbles are formed. High concentrations of OH- ions in the

solution created smaller nanobubbles with higher charge density values for a given

energy input than that at neutral pH solutions. Literature review showed that, under

wide range of solution pH values, zeta potential of nanobubbles was negative, and

negative value increased with increased solution pH values [14] [18] [19] [20]. It is

noted that upon increasing pH, negative zeta potential increased and reached a plateau

of approximately -110 mV at pH &10, and for acidic solutions with pH below 4.5,

zeta potential values were positive [14]. It is also noted that zeta potential values of
nanobubbles showed a sigmoidal behavior between pH 2 (+26 mV) and pH 8.5 (-28

mV) with isoelectric point (IEP) at pH 4.5 and highest negative zeta potential (-59

mV) at pH 10 [20]. When looking at size variation with solution pH, the same study

[20] showed that bubble size reached the maximum (720 nm) around an IEP value at

pH 4.5 where bubbles were practically uncharged (–5 mV). Finally, it was concluded

that the higher the amount of electrical charge on the bubble, it resulted in smaller

nanobubbles [20]. In addition, it is duly noted that bubbles created with solution pH

of 3 were much larger than those created with solution pH of 12, and effective

diameter was reported as 372 nm with a solution of pH 3 and 293.4 nm with a

solution of pH 12 [18]. Therefore, increasing the solution pH with high concentration

of OH- ions would increase the zeta potential. Moreover, this will increase hydrogen

bonds around the bubbles and will help to increase the stability of the bubble as well.

The experimental data from this research supported the above conclusion with the

highest magnitude of negative zeta potential of -27.3 V for a nano oxygen in a

solution pH of 10. Based on experimental results, the smallest zeta potential value

was obtained for a solution pH of 4. So, the reduction in negative zeta potential can

be easily attributed to high concentration of H + ions in the solution or a reduction of

OH- ion concentration. Figures 3b and 4b show that nanobubbles tend to be smaller in

size with increased solution pH values. In addition, they showed that in a solution

with a neutral pH and above, bubble size remained smaller in the nanosize range for 1

week. However, for a solution pH of 4, bubbles were much bigger in the microsize

range at the time of generation and very rapidly increased in size, and after weeks

after, the sizes cannot be analyzed anymore. Test results showed that stable bubbles
were generated under the neutral solution pH and for solution pH values above 7.

Even though nanobubbles in high pH NaOH solutions showed highly negative zeta

potential value at the time of generation, it rapidly reduced to values close to zeta

potential values of nanobubbles produced with neutral solution pH (Fig. 2a). Also, the

results revealed, nanobubbles in acidic solutions were difficult to generate and those

zeta potential values tend to be positive. This confirms that the surface charge of

nanobubbles is strongly related to the OH- ion concentration. Hence it can be

concluded that stable nanobubbles are generated under less acidic environments. This

can be explained as, when nanobubbles are formed in acidic solutions magnitude zeta

potential values are always low compared to the neutral or alkaline conditions. Those

bubbles with low zeta potential have higher possibility for bubble coalescence and,

therefore, generating unstable bubbles.

Figures 3b and 4b show that nanobubbles tend to be smaller in size with

increased solution pH values. In addition, they showed that in a solution with a

neutral pH and above, bubble size remained smaller in the nanosize range for 1 week.

However, for a solution pH of 4, bubbles were much bigger in the microsize range at

the time of generation and very rapidly increased in size, and after a week ZetaSizer

could not accurately measure the size (Fig. 2b).

Test results showed that stable bubbles were generated under the neutral

solution pH and for solution pH values above 7. Even though nanobubbles in high pH

NaOH solutions showed highly negative zeta potential value at the time of

generation, it rapidly reduced to values close to zeta potential values of nanobubbles

produced with neutral solution pH (Fig. 2a). Also, the results revealed, nanobubbles
in acidic solutions were difficult to generate and those zeta potential values tend to be

positive. This confirms that the surface charge of nanobubbles is strongly related to

the OH- ion concentration. Hence it can be concluded that stable nanobubbles are

generated under less acidic environments. This can be explained as, when

nanobubbles are formed in acidic solutions magnitude zeta potential values are

always low compared to the neutral or alkaline conditions. Those bubbles with low

zeta potential have higher possibility for bubble coalescence and, therefore, creating

unstable bubbles.

Figure 1 Relationship between Relationship between (a) zeta potential variation


with time and (b) bubble size variation with time, for oxygen nanobubble
prepared with different pH at 20C.
Figure 2. Relationship between
(a) zeta potential and the NaCl concentrations and (b) the bubble diameter and NaCl
concentrations, for ozone nanobubbles at 25C.

3. Size and Zeta Potential of Nanobubbles

The stable nanobubbles are smaller size bubbles having very negligible rising

velocity and contain high-magnitude zeta potential values to reduce the possibility of

bubble coalescence. To form high zeta potential values, it must contain high surface

charge density. As mentioned before, nano OH- bubble surface charge is related to the

ions or less hydrated and more polarized anions at the bubble gas–water interface. So

that, by adding surfactants, increasing pH, or using other methods, one can create a

favorable environment

OH- to generate ions/less hydrated and more polarized anions at the gas–water interface

and, hence, stable nanobubbles [12].

3.1. Particle Size, Other Properties, and Zeta Potential

Micro- and nano-bubbles (MNB) have diameter smaller than several tens of

micrometers. At this small size, bubbles are supposed to present different

physicochemical and fluid dynamic properties than ordinary macro-bubbles [21]

Among the physicochemical characteristics of MNB, there is the large

specific area and the high pressurization of gas inside the bubble, which confer to

these bubbles high gas dissolution capability [22]. Furthermore, MNB were reported

to have an electrically charged surface [14] [10]. It was also observed free-radical

generation with the micro- bubble collapse [14]. In addition, it is reported that air

micro-bubbles were pseudo-elastic and spherical in aqueous solutions [23].

Regarding the fluid dynamic properties, it was cited the low rising velocity under the
liquid phase and the reducing frictional resistance [24]. Besides these characteristics,

it is supposed that there are still some other unknown properties of MNB.

The mechanism of the acceleration of the physiological activity by the MNB

has not been elucidated yet. The main obstacle is to understand the characteristics of

the water containing MNB. One of the important properties of MNB is the electrical

charges on the bubble surface. The low electrical potential created by the charged

surfaces of MNB could be related to the physiological activation in living organisms

[21] The electrical potential of a particle in a colloidal system can be expressed by

the zeta-potential. The zeta-potential is defined as the electrical potential at the

slipping plane of a particle. The slipping plane is the boundary layer that divides the

double layer formed by the counter ions on the particle and the bulk solution [8].

The images obtained through the optical microscope with the scattering laser

showed particles in the range of few hundreds of nanometers. These particles moved

with the application of a voltage, indicating that they are electrically charged. The

zeta- potential of the water after the generation of micro- and nano- bubbles was

negative in all samples, but the values varied with the kind of gas. Higher absolute

zeta-potential were obtained in O2 and N2 MNB waters, while air, CO 2 and Xe

showed lower absolute values. The difference in zeta-potential could indicate

different bubble stability. The causes for the different zeta-potential values with

different kind of gases are still not clear, but the gas should play a role in the bubble

structure at nano-scale. Finally, the bubbling time seems to affect the zeta-potential

in the O2 NB water, although a more detailed investigation should be done to clarify

this point.
Chapter 3

METHODOLOGY

Nanobubbles are frequently generated in solutions by creating cavities. Cavitation is

caused by pressure reduction below the certain critical value. Based on the pressure re-

duction mechanism, cavitation mechanisms can be classified into four different types [25]

[26] [27].

 Hydrodynamic—variation in the pressure of liquid flux due to system geometry

[25] [26] [28]

 Acoustic—acoustic cavitation produced by applying ultrasound to liquids [29]

[30]

 Particle—passing high intensity light photons in liquids [31]

 Optical—short-pulsed lasers focused into low absorp- tion coefficient solutions

[32] [27]

Nanobubbles are usually hydrodynamically generated using the following methods

([24]

 Dissolve gases in liquids by compressing gas flows in liquids, then releasing those

mixtures through nano- sized nozzles to create nanobubbles.

 Inject low pressure gases into liquids to break gas into bubbles by focusing, fluid

oscillation, or mechanical vibration.


In this research, the hydrodynamic cavitation was used to generate nanobubble [25]

[26] [24] where BT-50FR micro- and nano-sized nozzle was used [12]. This gas–water

circulation method generates a flume of micro- and nano- bubbles in water. First, water is

pumped into the nozzle with an eccentricity to create a swirling effect. The swirling water

creates a vacuum at the outlet of the nozzle where the desired gas is injected at a controlled

rate. The gas introduced by vacuum into the swirling water will exit from the outlet as a

mixture of micro- and nanobubbles. The dynamic forces within the vortex will break the

injected air into smaller bubbles.

The main objective of this study was to investigate the long-term stability of

nanobubbles based on factors that influence the generation of nanobubbles. Therefore,

nanobubbles were generated under the following four different conditions and tested for their

size distributions and zeta potential values. Then, the results were analyzed.

1. Different type of Gases (test series IA)—air, oxygen, nitrogen, and ozone in

Deionized (DI) water.

2. Different pH levels (test series IIA)—DI water withNaOH and HCl to produce

solutions with different pH values (4, 7, and 10).

3. Different Salt Concentrations (test series IIB)—DI water with NaCl to prepare

different solution concentrations (0.001, 0.01, 0.1, and 1 M solutions).

4. Different Temperatures (test series IIC)[33]—DI water with 15C, 20C, and 30C

solution temperatures by heating or using a chiller.

Stability and reactivity of nanobubbles should depend on gases inside cavities.

Therefore, to investigate different gas bubble properties, nanobubbles were generated using
different gasses, namely, ozone, oxygen, air, and nitrogen. Bubbles were generated in DI

water where the electric conductivity of the solution was maintained at 0.3 mS/cm by adding

NaCl (test series IA) [12].

Materials and Solution Preparation. -Cyclodextrin (R-CD, GR) is purified by

recrystallization one time in 60% ethanol aqueous solution and then twice in water. The

purified R-CD was dried in a vacuum oven at 80 °C for ∼12 h.17 Sodium chloride (NaCl,

GR from BDH) impurities. Sodium hydroxide (NaOH, GR) is heated at ∼200 °C for ∼2 days

to remove organic and hydrochloric acid (HCl, 37% from Fisher) were used without further

purification. Water was purified with an inverse osmosis filtration (Nano Pure, Barnstead)

until its resistivity reached 18.2 MΩ‚cm at 20 °C and then filtered with a Millipore PTFE

0.45-µm hydrophilic filter. An R-CD aqueous solution (CR-CD ) 7.90 × 10-3 g/mL) was

prepared at ∼20 °C and ensure that it was saturated with gas. Note that the nanobubbles

bubbled by air with a flow rate of ∼15 mL/min for 2.5 h to can spontaneously form in

solution without the air purge. The air purge ensures that each solution studied has the same

starting point. To avoid a possible change of the ionic strength or pH, air was pumped

through a saturated NaOH aqueous solution to eliminate carbon dioxide (CO 2) and a

Millipore 0.22 µm Nylon filter to remove dust particles. The solution pH was adjusted by

addition of NaOH or HCl aqueous solution and monitored by a pH meter. For LLS

experiments, each solution was clarified with a Millipore 0.45-µm hydrophilic filter to

remove large dust particles.

Characterization of Nanobubble Suspensions. The size distribution and the number

density of bulk nanobubbles are quantified using a nanoparticle tracking analysis (NTA)

instrument). NTA tracks the Brownian motion of nanoparticles and is ideally suited for real-
time analysis of polydisperse systems ranging from 10 to 2000 nm in size and 107 to 109

particles per mL in concentration. It is superior to dynamic light scattering whose

measurements are based on the intensity of scattered light and is, thus, biased toward large

particles Two standard suspensions of solid latex nanospheres are to be used to verify the

accuracy and precision of the NTA system and to adjust the instrument settings accordingly,

prior to the analysis of nanobubble samples. The zeta potential of the nanobubbles should

also be included.

Physical and Chemical Analytical Techniques. Here, it describes the various

physical and chemical analytical techniques used to establish the evidence for the existence

of bulk nanobubbles in pure water and in aqueous ethanol solutions. Ethanol Separation from

Water−Ethanol Nanobubble Suspension Experiment. The aim of this process was to test what

happens when ethanol is removed from a nanobubble suspension produced by water−ethanol

mixing. Ethanol separation was carried out at 50 °C

FT-IR Analysis. FT-IR is a nondestructive, quantitative, and quick method for

identifying a wide range of chemical constituents and elucidating compound structures in

various forms in real-world samples according to the vibrational modes of their molecular

functional groups.40 FT-IR spectroscopic measurements were used here to investigate the

purity of nanobubble suspensions produced in pure water and in water−ethanol.


REFERENCES

[1] A. Villiers, Sur la transformation de la fécule en dextrine par le ferment butyrique",

Compt. Rend. Acad. Sci. (1891).

[2] G. Crini, Review: a history of cyclodextrins, Chem. Rev., 2014.

[3] T. Wimmer, No Title, Ullmann’s Encycl. Ind. Chem. (2012).

https://doi.org/10.1002/14356007.e08_e02.

[4] A. Valsesia, C. Desmet, I. Ojea-Jiménez, A. Oddo, R. Capomaccio, F. Rossi, P. Colpo,

Direct quantification of nanoparticle surface hydrophobicity, Commun. Chem. 1

(2018). https://doi.org/10.1038/s42004-018-0054-7.

[5] M. Chaplin, Nanobubbles (ultrafine bubbles)., (2017).

www1.lsbu.ac.uk/water/nanobubble.html.

[6] W. Soutter, What are nanobubbles?, (2013). https://www.azonano.com/article.aspx?

ArticleID=3151.

[7] I. Corporation, Ultrafine bubble generation technology, (2017).

www.idec.com/home/finebubble/bubble01%0A.html.

[8] Malvern Instruments Ltd., Detection and measurement of ultrafine bubbles, (2017).
www.malvern.com/en/%0Aindustry-applications/sample-type-form/nanobubbles.

[9] A. Azevedo, R. Etchepare, S. Calgaroto, J. Rubio, Aqueous dispersions of

nanobubbles: Generation, properties and features, Miner. Eng. 94 (2016) 29–37.

https://doi.org/10.1016/j.mineng.2016.05.001.

[10] M. Ushikubo, F.Y., Furukawa, T., Nakagawa, R., Enari, S. Makino, Y., Kawagoe, Y.,

Shiina, T., and Oshita, Evidence of the existence and the stability of nano-bubbles in

water, Elsevier. (2010). https://doi.org/10.1016/j.colsurfa.2010.03.005.

[11] J.L. Demangeat, Gas nanobubbles and aqueous nanostructures: The crucial role of

dynamization, 2015. https://doi.org/10.1016/j.homp.2015.02.001.

[12] J.N. Meegoda, S. Aluthgun Hewage, J.H. Batagoda, Stability of nanobubbles, Environ.

Eng. Sci. 35 (2018) 1216–1227. https://doi.org/10.1089/ees.2018.0203.

[13] J. Graciaa, A., Creux, P., Lachaise, Electrokinetics of bubbles. In Hubbard, A.T.,

Encycl. Surf. Colloid Sci. (2002) 829–845.

[14] T. M., Zeta potential of microbubbles in aqueous solutions: electrical properties of the

gas-water interface, J Phys Chem B. (n.d.). https://doi.org/0.1021/jp0445270.

[15] S.R. Rao, S.R. Rao, Hydrophobicity and Contact Angle, Surf. Chem. Froth Flotat.

(2004) 351–384. https://doi.org/10.1007/978-1-4757-4302-9_8.

[16] A. Agarwal, An experimental study of nanobubbles on hydrophobic surfaces, Dept.

Mech. Eng. Master of (2005) 127.

[17] G. Pan, B. Yang, Effect of surface hydrophobicity on the formation and stability of

oxygen nanobubbles, ChemPhysChem. 13 (2012) 2205–2212.


https://doi.org/10.1002/cphc.201100714.

[18] J.D. Kim, J.Y., Song, M.G., and Kim, Zeta potential of nanobubbles generated by

ultrasonication in aqueous alkyl polyglycoside solutions., J. Colloid Interface Sci. 223,

285. (2000).

[19] B.( Jia, W., Ren, S., and Hu, Effect of water chemistry on zeta potential of air

bubbles., Int. J. Electrochem. Sci. 8 (2013).

[20] J. Calgaroto, S., Wilberg, K.Q., and Rubio, On the nanobubbles interfacial properties

and future applications in flotation., Miner. Eng. (2014).

[21] Z. Serizawa, A., Inui, T., Yahiro, T., Kawara, Laminarization of micro-bubble

containing milky bubbly flow in a pipe, Proc. 3rd Eur. Japanese Two-Phase Flow Gr.

Meet. (2003) 21–27.

[22] R.M. Bredwell, M.D. and Worden, Mass-transfer properties of microbubbles, Exp.

Stud. Biotechnol. Prog. (1998) 31–38.

[23] N.A. Fan, X., Zhang, Z., Li, G., Rowson, Attachment of solid particles to air bubbles

in surfactant-free aqueous solutions, Chem. Eng. Sci. 59(13). (2004) 2639–2645.

[24] H. Tsuge, The latest technology on microbubbles and nanobubbles, C. Tokyo. (2007).

[25] L.U.O. Maoming, F.A.N., Daniel, T.A.O., Honaker, R., and Zhenfu, Nanobubble

generation and its application in froth flotation (part I): Nanobubble generation and its

effects on properties of microbubble and millimeter scale bubble solutions., Min. Sci.

Technol. (2010).

[26] Y. Agarwal, A., Ng, W.J., and Liu, Principle and ap- plications of microbubble and
nanobubble technology for water treatment, Chemosphere. (2011).

[27] R. Padilla-Martinez, J.P., Berrospe-Rodriguez, C., Aguilar, G., Ramirez-San-Juan,

J.C., and Ramos-Garcia, Optic cavitation with CW lasers: A review., Phys. Fluids. 26

(2014).

[28] J. Oliveira, H., Azevedo, A., and Rubio, Nanobubbles generation in a high-rate

hydrodynamic cavitation tube, Miner. Eng. (2018).

[29] M. Ashokkumar, The characterization of acoustic cavitation bubbles—An overview.,

Ultrason. Sonochem. 18 (2011).

[30] T.J. Ashokkumar, M., and Mason, Sonochemistry., Kirk-Othmer Encycl. Chem.

Technol. (2007).

[31] S. Poulain, S., Guenoun, G., Gart, S., Crowe, W., and Jung, Particle motion induced

by bubble cavitation, Phys. Rev. Lett. (2015).

[32] W. Lauterborn, OPTIC CAVITATION, . . Phys. Colloq. 40 (1979).

[33] A.J. Jadhav, M. Barigou, Bulk Nanobubbles or Not Nanobubbles: That is the

Question, Langmuir. 36 (2020) 1699–1708.

https://doi.org/10.1021/acs.langmuir.9b03532.

S-ar putea să vă placă și