Next Article in Journal
Novel Micro- and Nanocellulose-Based Delivery Systems for Liposoluble Compounds
Next Article in Special Issue
Molecular Dynamics Simulations on the Elastic Properties of Polypropylene Bionanocomposite Reinforced with Cellulose Nanofibrils
Previous Article in Journal
Enhanced Thermoelectric Performance of Polycrystalline Si0.8Ge0.2 Alloys through the Addition of Nanoscale Porosity
Previous Article in Special Issue
Revealing the Hemispherical Shielding Effect of SiO2@Ag Composite Nanospheres to Improve the Surface Enhanced Raman Scattering Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

From Bubbles to Nanobubbles

by
George Z. Kyzas
and
Athanasios C. Mitropoulos
*
Department of Chemistry, International Hellenic University, GR 65404 Kavala, Greece
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(10), 2592; https://doi.org/10.3390/nano11102592
Submission received: 11 August 2021 / Revised: 21 September 2021 / Accepted: 29 September 2021 / Published: 1 October 2021
(This article belongs to the Special Issue Polymer Based Nanocomposites: Experiment, Theory and Simulations)

Abstract

:
Nanobubbles are classified into surface and bulk. The main difference between them is that the former is immobile, whereas the latter is mobile. The existence of sNBs has already been proven by atomic force microscopy, but the existence of bNBs is still open to discussion; there are strong indications, however, of its existence. The longevity of NBs is a long-standing problem. Theories as to the stability of sNBs reside on their immobile nature, whereas for bNBs, the landscape is not clear at the moment. In this preliminary communication, we explore the possibility of stabilizing a bNB by Brownian motion. It is shown that a fractal walk under specific conditions may leave the size of the bubble invariant.

1. Introduction

If you would like to witness nature’s ability to create perfection, the best example is of a bubble. Nature not only knows how to make beautiful bubbles but also how to make them easily. In science, once the forces that are involved in the formation of a bubble were discovered, a whole new chapter of physical laws on surface thermodynamics emerged. To this end, some very notable scientists, such as Laplace, Plateau, Lord Rayleigh, Willard Gibbs, and many others, have left their mark on this area of science.
The first study was on soap bubbles [1] consisting of two surfaces, where each is tense or contractile. Balancing this pressure difference, Laplace obtained his famous law [2]:
Δ P = 4 γ R
where γ is the surface tension, and R is the radius of the bubble. Equation (1) shows that the smaller the bubble, the greater the pressure inside it, i.e., pressure and curvature rise and fall together.
Liquids have skin, but different liquids have different skin strengths. Although to blow a soap bubble is easy, to blow a water bubble is not. This is because the elastic strength of clean water is much less than that of soap solution. While water bubbles in air are not possible, air bubbles in water are a common phenomenon, and their stability has been a long-standing problem for more than 70 years [3,4,5,6,7].
In this study, we focus on nanobubbles (NBs), i.e., bubbles with sizes from 1 μm to 100 nm or less. According to diffusion theory, the lifetime of these bubbles ranges, respectively, from ms to μs. Therefore, they should not exist, but they do; not only do they exist, but they also persist almost indefinitely [8]. Figure 1a illustrates the whole bubble spectrum. According to ISO 20480-1-2017 [9], bulk NBs are also called ultrafine bubbles. Figure 1b shows this classification. The volume equivalent diameter deq is given by the following equation:
d e q = ( 6 V b u b b l e π ) 1 3
where Vbubble is the volume of the bubble, which also includes the bubble shell if the bubble is covered by a bubble shell.
Early explanations on why NBs do not obey diffusion theory have concluded that this is due to the immobilization of the interface by surface contaminants [10] or to the impediment of the bubble by a solid particle known as the Harvey-nucleus [11,12,13]. Recent developments include further explanations such as hard hydrogen-bonding [14,15], electrostatic repulsion [16], dynamic equilibrium [17], or contact line pinning [18]; meanwhile, the debate goes on [19,20,21,22,23,24,25,26,27,28].

2. Historical Definition of the Problem

In a water column that is under negative pressure, a bubble may spontaneously appear within the liquid [29]. The process is known as cavitation, and the magnitude of the negative pressure at which it takes place defines the tensile strength of the liquid. In the case of homogeneous nucleation, the total energy required to form such a bubble is given by:
Δ E = 4 π R 2 γ 4 3 π R 3 Δ p = 4 3 π R 2 γ
As the size of the nucleus grows, it reaches a critical (maximum) radius Rc just before the liquid raptures. At this point, the energy has a maximum value, ΔEmax, which is formulated by Gibbs as:
Δ E max = 16 π γ 3 3 Δ p c ,
where Δpc is the pressure difference across the interface at the critical radius of the bubble.
In 1850, Berthelot [30] measured the tensile strength of purified water to about 50 atm. On the other hand, the theoretical calculations of Frenkel [31] predicted a tension of 100 times the experimental one. Cavitation may, thus, be considered a precursor of NBs, for which the theory also fails to explain.
Let us denote with c the gas concentration far away from the bubble and with cs the gas concentration at the bubble–liquid interface. In an undersaturated solution, c < cs, the NBs should dissolve away if the ambient pressure p is sufficiently high. In an oversaturated solution, c > cs, the NBs should grow, then rise and burst. In an even solution, c = cs, the system is unstable, and the slightest disturbance will cause the bubble to either expand or to dissolve. Nevertheless, experience defies this prediction.
The mass transfer of gas in the liquid is governed by the diffusion equation [3]:
c t + d R d t ( R r ) 2 c t = D r 2 r ( r 2 c r ) ,  
where c(r,t) is the gas concentration in the liquid at time t and distance r from the center of a bubble of radius Ro at t = 0, and D is the mass diffusivity. Since a bubble in a liquid has only one surface, the Laplace equation reduces by a factor of 2:
Δ p = 2 γ R .
For a microbubble of R = 1 μm in water of γ = 72.8 mΝ/m, Δp = 1.5 bar, whereas for a nanobubble of 100 nm, Δp = 15 bar. According to Henry’s law:
p = K H c s ,  
where p is the pressure of the gas in the bubble, and KH is the Henry constant. By neglecting the second term on the left-hand side of Equation (5), Epstein and Plesset [4] found a solution that takes on the form:
ψ 2 = 1 ± X 2   with   Ψ = R R o   and   X 2 = 2 D c s ρ ( c c s 1 ) t ,  
where the (+) sign is for oversaturation (bubble expansion), the (−) sign is for undersaturation (bubble shrinkage), and ρ is the gas density. In the case of an air/water bubble at 295 K, cs/ρ = 0.02 and D = 2 × 10−9 m2/s. Figure 2 shows the growth and shrinkage of an NB of Ro = 100 nm, and Table 1 records the fate of different size bubbles according to this theory.
A perturbation as small as 0.01% on the equilibrium concentration will cause a bubble of radius 1 μm to dissolve in 2 min or to grow, without bounds, 10 times its original size in 3.4 h.

3. Explanations for NB Longevity

Fox and Herzfeld [10] have suggested that organic impurities (e.g., fatty acids) are adsorbed on the surface of the bubble and, as it contracts, will cover it completely with an organic skin that will slow down its contraction, although it will not reduce it to zero. A thin square sheet of area Ao under a uniform side force F will cause a relative change of that area by the following calculation [32]:
A A o = Y δ 1 ν F ,  
where Y is the Young modulus, δ is the thickness of the film, and ν is Poisson’s constant. For a tearing strength of 50 mN/m and δ = 15 Å, one gets a strength of 33 bar. However, later on, Herzfeld by himself [33], based on experimental data [34], abandoned the rigid skin hypothesis.
Instead of an organic skin, Akulichev [35] proposed an ionic skin based on the fact that air/water bubbles possess an electrical charge [36]. Hydrophobic ions such as Cl will migrate to the surface of the bubble, whereas hydrophilic ones such as OH will not.
Sirotyuk [37] extended Fox and Herzfeld theory by suggesting stabilization due to a film of surface-active agents. In the case of air/water NBs, the polar head will bond to the water, and the tail will dangle in the air. This skin now has greater elasticity and will help the bubble to stabilize. Only traces of surface-active agents are needed to save them.
Harvey and coworkers [11] studied bubble formation in animals. They considered a solid impurity with a crevice containing gas. When the receding angle θr is small, by increasing the pressure Δp across the gas–liquid interface, the radius of curvature decreases until r = R. Now, the bubble is unstable, and the slightest increase in Δp will cause the bubble to grow and the radius of curvature to increase. When the contact angle θ = θr, the bubble will move up the side of the cavity until the buoyant force becomes greater than 2πRγ and burst. Figure 3 shows this course of events, where: r1 > r2 > r = R < r3 < r4. If, on the other hand, θr is large enough, when θr = θ, the bubble will grow unpinned and eventually will move out of the cavity, rise, and burst. In both cases, a mass of gas may be left behind.
In order for the gas bubble to survive, the interface must be concave toward the liquid. This can be realized if the crevice is a cone with an acute apical angle, ω, so that: ca > (π + ω)/2, where ca is the advancing angle. Here, gas is creeping up the side of the cone until θ = θr. The condition for equilibrium is now Δp = −2γ/R, where the negative sign indicates a concave curvature. Figure 4 illustrates this mechanism.
Recent advancements on NBs classify them to bulk and surface, with the main difference being the lack of a three-phase contact line for the former compared to the latter [38]. As a result, bulk NBs are mobile, whereas the surface is not, and, again, the radius of the curvature of bulk NBs is much smaller than that of surface NBs. Furthermore, the existence of sNBs has already been proven by atomic force microscopy (AFM) [39,40,41], while the existence of bNBs is indicative but not secured [14,42,43,44,45]. This doubt is due to the lack of an appropriate experimental method; dynamic light scattering (DLS), which is the prime experimental technique for detecting bNBs, cannot distinguish between bubbles, droplets, or particles [46]. However, the pursue of new techniques in determining both surface and bulk NBs is of vital importance for the advancement of NB technologies and their applications. To this end, Pan et al. [47] have investigated O2 bubbles of >25 nm at a diatomite particle in situ with synchrotron-based scanning transmission soft X-ray microscopy (STXM). They strongly indicate that in situ studies provide useful information on material preparation, phase equilibrium, nucleation kinetics, and chemical composition in the confined space.
The maximum work Wc required for the formation of a surface bubble [48] of critical size is given by an equation similar to Equation (4), but with a factor Φ:
W c = 16 π γ 3 3 Δ p c Φ   with   Φ = 1 4 ( 2 + cos θ c ) × ( 1 cos θ c ) 2 ,
where θc is the critical contact angle from the side of gas. Based on this geometry, Brenner and Lohse [17] have introduced a dynamic equilibrium mechanism for sNB stabilization, where the gas out-flux Jout is compensated by gas in-flux Jin at the contact line.
J o u t = π R D ( 1 c c s ) = 2 π s D R tan θ ,
where s > 0 is the attraction potential of gas from the hydrophobic wall, θ is the contact angle, and R is now the radius of the spherical cap circle. However, this model does not fit the experimental results, and some years later, Lohse and Zhang [49] suggested that pinning and gas oversaturation may explain the stability of sNBs. When the bubble shrinks, the radius of the curvature increases; therefore, Δp decreases and eventually becomes too weak to press out the bubble against the oversaturation. Figure 5a simulates this mechanism; for comparison, Figure 5b shows the case of free-standing sNBs.
It is noted, however, that experiments show that sNBs survive in open systems and undersaturated environments too. Based on the AFM images, Qian et al. [50] have argued that pinning, although good for the stabilization process, is not enough; it has to be accompanied by another effect to explain the observed stability. They suggested that a kinetic barrier may prevent the rapid transfer of gas across the interface of the sNB. Tan et al. proposed that in an undersaturated solution, the sNB would be stabilized by both pinning and attractive hydrophobic forces. By dividing the bubble into slices of thickness (dz; see Figure 5b), they developed the following equation for a localized concentration buildup, c(z):
c ( z ) = c exp ( ϕ o e z λ k B T )
where ϕοe−z/λ is the short-ranged potential, which is attractive when ϕo < 0 and repulsive when ϕo > 0; λ = 1 nm is the interaction distance. Equation (12) predicts either a localized oversaturation next to a hydrophobic substrate or a localized undersaturation next to a hydrophilic solid.
Based on IR spectra for O-H, Ohgaki et al. [14] have suggested that hydrogen bonding on the interface of a bNB is similar to that of gas hydrates. This hard hydrogen bonding helps the interface to resist the diffusivity of the gas from the bNB and to maintain a kinetic balance against the internal pressure.
Experimental results have confirmed that NBs in water are negatively charged [16]. The zeta potential is about ζ = −40 mV in pure water (pH = 7). However, different values for different water purifications have been observed too. For instance, air NBs in deionized water have ζ = −65 mV, whereas in distilled water, ζ = −35 mV. Electrostatic repulsion may be added to the Laplace equation in a way that:
Δ P = 2 γ R ε ζ 2 R 2 ,  
where ε is the dielectric constant. However, the second term of the RHS of Equation (13) is small to balance the Laplace pressure. Zang et al. and others [51,52] have proposed a correction for surface tension at higher pressures:
γ = γ 0 ( 1 ρ ρ w ) 4 ,  
where ρw = 1000 kg/m3 is the water density. In the case of an air bubble of R = 100 nm and ρ = 20 kg/m3, γ = 67 mN/m. However, even this correction does not qualify Equation (13) as a solution to the problem. Nevertheless, the negative charge of bNB surfaces will prevent them from coalescing and, again, will form an electric double layer [53]. These electrokinetic properties of bNBs may help them to survive for a longer time. Moreover, the many bNBs presented in the solution may also help their stabilization because the large concentrations of bNBs can supply gas to the liquid, retarding their dissolution [54]. Figure 6 illustrates the electrical double layer for bNBs.

4. Discussion

The NB survival problem is very old and very complex, one reason being that the phenomenon is not static but dynamic. Epstein and Plesset, in their seminal paper on the stability of gas bubbles in liquid–gas solutions, have made several assumptions. For instance, the transport term of the diffusion equation has been omitted, as well as the vapor pressure (only the gas pressure is considered). Again, the content of the bubble, the temperature, and the pressure inside it are assumed to be homogeneous and uniform, and, of course, the shape of the bubble is understood to be spherical. In the solution, the pressure far away from the bubble, the liquid density, and the viscosity are taken to be constant and uniform and the liquid to be incompressible. Although these assumptions are shown to be valid for bubbles as small as 1 μm [55], they may not be justified in the case of smaller ones.
Bubbles in a solution are subject to different kinds of motion; buoyancy rise and Brownian motion are of interest here. The terminal velocity υb due to buoyancy, balanced by the Rybczynski approximation, equals to [56]:
υ b = g ρ w 3 η R 2 ,
where g is the acceleration of gravity and η is the water viscosity. On the other hand, the mean velocity υB due to Brownian motion is given by Einstein’s formula:
[ x 2 t 2 ] 1 2 = 2 D t = R G T 3 π η Ν A R t = υ B ,  
where <x2> is the average particle displacement, RG is the gas constant, NA is Avogadro’s number, and T is the absolute temperature. Apparently, as t approaches 0, Equation (16) diverges and, therefore, does not represent the real velocity [57,58]. However, the smaller the bubble, the greater the velocity. Figure 7 compares the two given velocities for air/water bubbles of different radii.
The cross size of the bubbles is ~1 μm. At this point, the average velocity is about 1 μm/s. For smaller bubbles, the Brownian motion dominates, and for larger ones, buoyancy prevails. In a saturated solution, a disturbance in the concentration of c–cs = 0.01% will require 4 s for a 100 nm bubble to double its radius. At the same time, this doubled size bubble will be displaced ~6 μm away from its original position; that is, 60 times its original radius. Although the gas concentration in the bulk is taken to be spherically symmetric and the density in the bubble uniform, the motion of the bubble may put them in doubt.
Nour nanoparticle tracking technique combines Brownian motion and light scattering to determine the bNBs’ diffusion constant and then to predict their spherical hydrodynamic diameter [59]. The moving bubble may not be spherical. Molecular dynamic simulations also point to the same conclusion [38]. As a result, the density of the gas within the bubble may not be uniform. Under Brownian motion, a bNB in a liquid, besides its inertia, will also be influenced by the inertia of the surrounding liquid. At short time scales (e.g., ns), such motion will also cause long-lived vortices in the liquid [57].
Another important physicochemical difference [60] between bulk and surface NBs is the Knudsen number, Kn. For a bNB, the Kn is always less than one: Kn = λ/R, where λ is the mean free path of the gas molecule. Therefore, the collision of gas molecules between each other dominates. The opposite behavior is common for sNBs, Kn = λ/h > 1, where h is the height of the spherical cap [61]. In this case, the collision of gas molecules with the bubble walls prevails. By excluding contamination as a universal mechanism for NB longevity of both bulk and surface, the solution to the problem must reside on their main difference, that is, the mobility of the former and immobility of the latter. Zou et al. [62] have studied bNBs in deionized water. They concluded that the bubbles remained stable for about two months due to Brownian motion. However, it must be noted that although several studies [63,64] have indicated that bNBs experience Brownian motion, the determination of the exact mechanism is still intriguing.
In an oversaturated solution, when the bubble is stationary, the size of the bubble increases and the Laplace pressure decreases, causing a reduction in gas concentration in the vicinity of the bubble. As a result, more gas diffuses from the solution into the bubble, thus establishing a positive feedback loop. The opposite takes place when starting from an undersaturated solution. If we now assume that the solution is in an even situation, the slightest perturbation, over or under, will initiate one of the aforementioned processes, and the bubble either will grow and burst or shrink and dissolve. However, in the domain of the solution, since such a perturbation in another domain must be an opposite one, this thus establishes a long wave of perturbations that cover up all the extend of the solution. If the bubble is now from an oversaturated domain displaced by Brownian motion, which is known to be persistent to an undersaturated domain and so on, it has a chance to survive. Figure 8 illustrates a schematic representation of this perception.
In Equation (8), by assuming the growth of the bubble and then shrinkage of the expanded bubble under the same but opposite conditions and for equal time, we get:
R + 2 R o 2 = + 2 D c s ρ ( c c s 1 ) t R 2 R + 2 = 2 D c s ρ ( c c s 1 ) t } R o 2 R + 2 = R 2 R + 2 R o 2 = R 2
where R+ and R are the radii of the expanded and contracted bubbles, respectively. As a result, the average size of the bubble remains invariant. The displacements of the bubble in Figure 8a have a pattern shown in Figure 8b:
x + 2 + x 2 = x 2 ,
where <x+2> and <x2> are the average displacements from over- to under-saturation. Hence, the generator of this pattern is made by two equal intervals of an angle of 90° alternating between the right and left of the teragon. This is a Peano curve with a fractal dimension H = 2 [65], thus connecting classical theory with the fractal character of Brownian motion [66,67].

5. Conclusions

Although there is an intense debate on the fundamentals of NB survival, there has not been equal hesitation from the industry on their usefulness. Applications to biology, water cleaning, the food industry, agriculture and hydroponics, and drug delivery ultrasonography are only some of the examples of using NBs successfully [68,69,70,71,72,73]. The Fine Bubble Industries Association (FBIA) and the Wall Street Journal have shown the business growth of NBs to have been from USD 20 million in 2010 to USD 10 billion in 2020. In the EU, the business is expected to grow from EUR 72 million in 2020 to EUR 145 million in 2030, where 52% of the market will be in the water treatment sector [46,74,75]. In this introductory note, we hope to ignite a discussion of new theories and radical hypotheses explaining how NBs avoid the Laplace catastrophe. Intuitively, it is understood that bubbles less than 1 μm in size may survive much longer than larger ones. At this size, diffusive Brownian motion crosses with buoyancy. Once a bubble is in the former region, it has a chance to survive for a longer time, while in the latter, it has not.

Author Contributions

Writing—original draft preparation, G.Z.K. and A.C.M.; writing—review and editing, G.Z.K. and A.C.M.; supervision, G.Z.K. and A.C.M. Both authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the project entitled “Development of Nanotechnology-Enabled Next-Generation Membranes and their Applications in Low-Energy, Zero-Liquid Discharge Desalination Membrane Systems”/NAMED (T2ΔΓΕ-0597), which is greatly acknowledged.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Boys, C.V. Soap Bubbles, Society for Promoting Christian Knowledge; Outlook Verlag: London, UK, 1916. [Google Scholar]
  2. Adamson, A.W.; Gast, A.P. Physical Chemistry of Surfaces, 6th ed.; John Wiley & Sons, Inc.: New York, NY, USA, 1997. [Google Scholar]
  3. Brennen, C.E. Cavitation and Bubble Dynamics; Oxford University Press: Oxford, UK, 1995. [Google Scholar]
  4. Epstein, P.S.; Plesset, M.S. On the stability of gas bubbles in liquid-gas solutions. J. Chem. Phys. 1950, 18, 1505–1509. [Google Scholar] [CrossRef] [Green Version]
  5. Plesset, M.S.; Sadhal, S.S. On the stability of gas bubbles in liquid-gas solutions. Appl. Sci. Res. 1982, 38, 133–141. [Google Scholar] [CrossRef] [Green Version]
  6. Mori, Y.; Hijikata, K.; Nagatani, T. Fundamental study of bubble dissolution in liquid. Int. J. Heat Mass Transf. 1977, 20, 41–50. [Google Scholar] [CrossRef]
  7. Cha, Y.S. On the equilibfium of cavitation nuclei in liquid-gas solutions. J. Fluids Eng. 1981, 103, 425–430. [Google Scholar] [CrossRef]
  8. Ball, P. Nanobubbles are not a superficial matter. ChemPhysChem 2012, 13, 2173–2177. [Google Scholar] [CrossRef] [PubMed]
  9. ISO 20480-1:2017. Fine Bubble Technology-General Principles for Usage and Measurement of Fine Bubbles–Part 1: Terminology. Available online: https://www.iso.org/standard/68187.html (accessed on 10 September 2021).
  10. Fox, F.E.; Herzfeld, K.F. Gas bubbles with organic skin as cavitation nuclei. J. Acoust. Soc. Am. 1954, 26, 984–989. [Google Scholar] [CrossRef]
  11. Harvey, E.N.; Barnes, D.K.; McElroy, W.D.; Whiteley, A.H.; Pease, D.C.; Cooper, K.W. Bubble formation in animals. I. Physical factors. J. Cell. Comp. Physiol. 1944, 24, 1–22. [Google Scholar] [CrossRef]
  12. Crum, L.A. Nucleation and stabilization of microbubbles in liquids. Flow Turbul. Combust. 1982, 38, 101–115. [Google Scholar] [CrossRef]
  13. Jones, S.; Evans, G.; Galvin, K. Bubble nucleation from gas cavities—A review. Adv. Colloid Interface Sci. 1999, 80, 27–50. [Google Scholar] [CrossRef]
  14. Ohgaki, K.; Khanh, N.Q.; Joden, Y.; Tsuji, A.; Nakagawa, T. Physicochemical approach to nanobubble solutions. Chem. Eng. Sci. 2010, 65, 1296–1300. [Google Scholar] [CrossRef]
  15. Nakashima, S.; Spiers, C.J.; Mercury, L.; Fenter, P.A.; Hochella, M.F., Jr. Physicochemistry of Water in Geological and Biological Systems—Structures and Properties of Thin Aqueous Films; Universal Academy Press Inc.: Tokyo, Japan, 2004; pp. 2–5. [Google Scholar]
  16. Takahashi, M. Potential of microbubbles in aqueous solutions: Electrical properties of the gas-water interface. J. Phys. Chem. B 2005, 109, 21858–21864. [Google Scholar] [CrossRef]
  17. Brenner, M.P.; Lohse, D. Dynamic equilibrium mechanism for surface nanobubble stabilization. Phys. Rev. Lett. 2008, 101, 214505. [Google Scholar] [CrossRef] [Green Version]
  18. Liu, Y.; Zhang, X. Nanobubble stability induced by contact line pinning. J. Chem. Phys. 2013, 138, 014706. [Google Scholar] [CrossRef] [PubMed]
  19. Ducker, W.A. Contact Angle and Stability of Interfacial Nanobubbles. Langmuir 2009, 25, 8907–8910. [Google Scholar] [CrossRef] [PubMed]
  20. Weijs, J.H.; Snoeijer, J.H.; Lohse, D. Formation of surface nanobubbles and the universality of their contact angles, A molecular dynamics approach. Phys. Rev. Lett. 2012, 108, 104501. [Google Scholar] [CrossRef] [Green Version]
  21. Kyzas, G.Z.; Favvas, E.P.; Kostoglou, M.; Mitropoulos, A.C. Effect of agitation on batch adsorption process facilitated by using nanobubbles. Colloids Surf. A 2020, 607, 125440. [Google Scholar] [CrossRef]
  22. Weijs, J.; Lohse, D. Why Surface Nanobubbles Live for Hours. Phys. Rev. Lett. 2013, 110, 054501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhang, X.; Chan, D.; Wang, D.; Maeda, N. Stability of Interfacial Nanobubbles. Langmuir 2013, 29, 1017–1023. [Google Scholar] [CrossRef]
  24. Chan, C.U.; Arora, M.; Ohl, C.-D. Coalescence, Growth, and Stability of Surface-Attached Nanobubbles. Langmuir 2015, 31, 7041–7046. [Google Scholar] [CrossRef]
  25. Eshibri, M.; Qian, J.; Jehannin, M.; Craig, V.S.J. A History of Nanobubbles. Langmuir 2016, 32, 11086–11100. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, C.; Li, J.; Zhang, X. The existence and stability of bulk nanobubbles: A long-standing dispute on the experimentally observed mesoscopic inhomogeneities in aqueous solutions. Commun. Theor. Phys. 2020, 72, 037601. [Google Scholar] [CrossRef]
  27. Ayodele, A.T.; Valizadeh, A.; Adabi, M.; Esnaashari, S.S.; Madani, F.; Khosravani, M. Ultrasound nanobubbles and their applications as theranostic agents in cancer therapy: A review. Biointerface Res. Appl. Chem. 2017, 7, 2253–2262. [Google Scholar]
  28. Michailidi, E.D.; Bomis, G.; Varoutoglou, A.; Kyzas, G.; Mitrikas, G.; Mitropoulos, A.C.; Efthimiadou, E.K.; Favvas, E.P. Bulk nanobubbles: Production and investigation of their formation/stability mechanism. J. Colloid Interface Sci. 2019, 564, 371–380. [Google Scholar] [CrossRef] [PubMed]
  29. Maris, H.; Balibar, S. Negative Pressures and Cavitation Liquid Helium. Phys. Today 2000, 53, 29–34. [Google Scholar] [CrossRef] [Green Version]
  30. Berthelot, M. Sur quelques phenomenes de dilation forcee de liquides. Ann. Chim. Phys. 1850, 30, 232–237. [Google Scholar]
  31. Frenkel, J. Kinetic Theory of Liquids; The Clarendon Press: Oxford, UK, 1946. [Google Scholar]
  32. Garabedian, C.A.; Love, A.E.H. The mathematical theory of elasticity. Am. Math. Mon. 1928, 35, 196. [Google Scholar] [CrossRef]
  33. Herzfeld, K.F. Proceedings of First Symposium on Naval Hydrodynamics; Sherman, F.S., Ed.; National Academy of Sciences: Washington, DC, USA, 1957; pp. 319–320. [Google Scholar]
  34. Strasberg, M. Onset of ultrasonic cavitation in tap watet. J. Acoust. Soc. Am. 1959, 31, 163–176. [Google Scholar] [CrossRef]
  35. Akulichev, V.A. Hydration of ions and the cavitation resistance of water. Sov. Phys. Acoust. 1966, 12, 144–149. [Google Scholar]
  36. Alty, T. The origin of the electrical charge on small particles in water. Proc. R. Soc. London. Ser. A Math. Phys. Sci. 1926, 112, 235–251. [Google Scholar] [CrossRef]
  37. Sirotyuk, M.G. Stabflization of gas bubbles in water. Sov. Phys. Acoust. 1970, 16, 237–240. [Google Scholar]
  38. Li, C.; Zhang, A.M.; Wang, S.; Cui, P. Formation and coalescence of nanobubbles under controlled gas concentration and species. AIP Adv. 2018, 8, 015104. [Google Scholar] [CrossRef]
  39. Ishida, N.; Inoue, T.; Miyahara, M.; Higashitani, K. Nano bubbles on a hydrophobic surface in water observed by tapping-mode atomic force microscopy. Langmuir 2000, 16, 6377–6380. [Google Scholar] [CrossRef]
  40. Lou, S.-T.; Ouyang, Z.-Q.; Zhang, Y.; Li, X.-J.; Hu, J.; Li, M.-Q.; Yang, F.-J. Nanobubbles on solid surface imaged by atomic force microscopy. J. Vac. Sci. Technol. B Microelectron. Nanometer Struct. 2000, 18, 2573. [Google Scholar] [CrossRef]
  41. Zhang, X.H.; Khan, A.; Ducker, W.A. A nanoscale gas state. Phys. Rev. Lett. 2007, 98, 136101. [Google Scholar] [CrossRef] [Green Version]
  42. Sugano, K.; Miyoshi, Y.; Inazato, S. Study of Ultrafine Bubble Stabilization by Organic Material Adhesion. Jpn. J. Multiph. FLOW 2017, 31, 299–306. [Google Scholar] [CrossRef] [Green Version]
  43. Yasui, K.; Tuziuti, T.; Kanematsu, W. Mysteries of bulk nanobubbles (ultrafine bubbles); stability and radical formation. Ultrason. Sonochemistry 2018, 48, 259–266. [Google Scholar] [CrossRef]
  44. Nirmalkar, N.; Pacek, A.W.; Barigou, M. On the existence and stability of bulk nanobubbles. Langmuir 2018, 34, 10964–10973. [Google Scholar] [CrossRef] [PubMed]
  45. Tan, B.H.; An, H.; Ohl, C.-D. How Bulk Nanobubbles Might Survive. Phys. Rev. Lett. 2020, 124, 134503. [Google Scholar] [CrossRef]
  46. Tan, B.H.; An, H.; Ohl, C.-D. Stability of surface and bulk nanobubbles. Curr. Opin. Colloid Interface Sci. 2021, 53, 101428. [Google Scholar] [CrossRef]
  47. Pan, G.; He, G.; Zhang, M.; Zhou, Q.; Tyliszczak, T.; Tai, R.; Guo, J.; Bi, L.; Wang, L.; Zhang, H. Nanobubbles at hydrophilic particle−water interfaces. Langmuir 2016, 32, 11133–11137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Fisher, J.C. The fracture of liquids. J. Appl. Phys. 1948, 19, 1062–1067. [Google Scholar] [CrossRef]
  49. Loshe, D.; Zhang, X. Pinning and gas oversaturation imply stable single surface nanobubbles. Phys. Rev. E 2015, 91, 031003. [Google Scholar]
  50. Qian, J.; Craig, V.S.J.; Jehannin, M. Long-term stability of surface nanobubbles in undersaturated aqueous solution. Langmuir 2019, 35, 718–728. [Google Scholar] [CrossRef] [PubMed]
  51. Zhang, L.; Chen, H.; Li, Z.; Fang, H.; Hu, J. Long lifetime of nanobubbles due to high inner density. Sci. China Ser. G-Phys. Mech. Astron. 2008, 51, 219–224. [Google Scholar] [CrossRef]
  52. Ulatowski, K.; Sobieszuk, P.; Mróz, A.; Ciach, T. Stability of nanobubbles generated in water using porous membrane system. Chem. Eng. Process. Process. Intensif. 2018, 136, 62–71. [Google Scholar] [CrossRef]
  53. Jia, W.; Ren, S.; Hu, B. Effect of water chemistry on zeta potential of air bubbles. Int. J. Electrochem. Sci. 2013, 8, 5828–5837. [Google Scholar]
  54. Weijs, J.H.; Seddon, J.R.T.; Lohse, D. Diffusive Shielding Stabilizes Bulk Nanobubble Clusters. ChemPhysChem 2012, 13, 2197–2204. [Google Scholar] [CrossRef] [Green Version]
  55. Duncan, P.B.; Needham, D. Test of the Epstein−Plesset Model for Gas Microparticle Dissolution in Aqueous Media: Effect of Surface Tension and Gas Undersaturation in Solution. Langmuir 2004, 20, 2567–2578. [Google Scholar] [CrossRef]
  56. Lamb, H. Hydrodynamics, 6th ed.; Cambridge University Press: London, UK, 1932. [Google Scholar]
  57. Li, T.; Raizen, M.G. Brownian motion at short time scales. Ann. Phys. 2013, 525, 281–295. [Google Scholar] [CrossRef] [Green Version]
  58. Chicea, D. Coherent light scattering on nanofluids: Computer simulation results. Appl. Opt. 2008, 47, 1434–1442. [Google Scholar] [CrossRef] [PubMed]
  59. Ultrafine Bubbles Recorded by NanoSight. 2019. Available online: www.acniti.com (accessed on 10 August 2021).
  60. Seddon, J.R.T.; Lohse, D.; Ducker, W.A.; Craig, V.S.J. A Deliberation on Nanobubbles at Surfaces and in Bulk. ChemPhysChem 2012, 13, 2179–2187. [Google Scholar] [CrossRef]
  61. Seddon, J.R.T.; Zandvliet, H.J.; Lohse, D. Knudsen gas provides nanobubble stability. Phys. Rev. Lett. 2011, 107, 116101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Zhou, Y.; Han, Z.; He, C.; Feng, Q.; Wang, K.; Wang, Y.; Luo, N.; Dodbiba, G.; Wei, Y.; Otsuki, A.; et al. Long-Term Stability of Different Kinds of Gas Nanobubbles in Deionized and Salt Water. Materials 2021, 14, 1808. [Google Scholar] [CrossRef] [PubMed]
  63. Nirmalkar, N.; Pacek, A.; Barigou, M. Interpreting the interfacial and colloidal stability of bulk nanobubbles. Soft Matter 2018, 14, 9643–9656. [Google Scholar] [CrossRef] [Green Version]
  64. Oh, S.H.; Kim, J.-M. Generation and Stability of Bulk Nanobubbles. Langmuir 2017, 33, 3818–3823. [Google Scholar] [CrossRef] [PubMed]
  65. Mandelbrot, B.B. The Fractal Geometry of Nature; Freeman Co.: New York, NY, USA, 1982. [Google Scholar]
  66. Vicsek, T.; Gould, H. Fractal Growth Phenomena. Comput. Phys. 1989, 3, 108. [Google Scholar] [CrossRef]
  67. Saberi, A.A. Fractal structure of a three-dimensional Brownian motion on an attractive plane. Phys. Rev. E 2011, 84, 021113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Mitropoulos, A.C.; Bomis, G. Device for Generating and Handling Nanobubbles. European Patent EP2995369A1, 2016. [Google Scholar]
  69. Favvas, E.P.; Kyzas, G.Z.; Efthimiadou, E.K.; Mitropoulos, A.K. Bulk nanobubbles, generation methods and potential applications. Curr. Opin. Colloid Interf. Sci. 2021, 54, 101455. [Google Scholar] [CrossRef]
  70. Kyzas, G.Z.; Bomis, G.; Kosheleva, R.I.; Efthimiadou, E.K.; Favvas, E.P.; Kostoglou, M.; Mitropoulos, A.C. Nanobubbles effect on heavy metal ions adsorption by activated carbon. Chem. Eng. J. 2019, 356, 91–97. [Google Scholar]
  71. Agarwal, A.; Ng, W.J.; Liu, Y. Principle and applications of microbubble and nanobubble technology for water treatment. Chemosphere 2011, 84, 1175–1180. [Google Scholar] [CrossRef] [PubMed]
  72. Ebina, K.; Shi, K.; Hirao, M.; Hashimoto, J.; Kawato, Y.; Kaneshiro, S.; Morimoto, T.; Koizumi, K.; Yoshikawa, H. Oxygen and Air Nanobubble Water Solution Promote the Growth of Plants, Fishes, and Mice. PLoS ONE 2013, 8, e65339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Liu, S.; Kawagoe, Y.; Makino, Y.; Oshita, S. Effects of nanobubbles on the physicochemical properties of water: The basis for peculiar properties of water containing nanobubbles. Chem. Eng. Sci. 2013, 93, 250–256. [Google Scholar] [CrossRef]
  74. Fine Bubble Industries Association. Available online: fbia.or.jp (accessed on 9 August 2021).
  75. Koltsov, D.K. Fine Bubble Technology in the EU; BREC Solutions Ltd.: Glasgow, UK, 2016. [Google Scholar]
Figure 1. (a) The spectrum of bubble sizes according to their stability. (b) Bubble diameter classification according to ISO 20480-1-2017 [9].
Figure 1. (a) The spectrum of bubble sizes according to their stability. (b) Bubble diameter classification according to ISO 20480-1-2017 [9].
Nanomaterials 11 02592 g001
Figure 2. Shrinkage or growth of a bubble of radius 100 nm at various concentrations.
Figure 2. Shrinkage or growth of a bubble of radius 100 nm at various concentrations.
Nanomaterials 11 02592 g002
Figure 3. A bubble impended in a crevice. The course of events Δp increases, where r is the radius of curvature, θr is the receding angle (small), and R is the pinning half distance between Points A and B. When r = R, the bubble is hemispherical.
Figure 3. A bubble impended in a crevice. The course of events Δp increases, where r is the radius of curvature, θr is the receding angle (small), and R is the pinning half distance between Points A and B. When r = R, the bubble is hemispherical.
Nanomaterials 11 02592 g003
Figure 4. A bubble impended in a cone crevice of a very small apical angle, ω, and θ, the equilibrium contact angle, with θa as the advancing angle and θr the receding angle. Notice that the meniscus is now concave.
Figure 4. A bubble impended in a cone crevice of a very small apical angle, ω, and θ, the equilibrium contact angle, with θa as the advancing angle and θr the receding angle. Notice that the meniscus is now concave.
Nanomaterials 11 02592 g004
Figure 5. (a) Stabilization of an NB pinned on a surface. As the height and the contact angle decrease, the radius of the curvature increases and the pressure across the interface deflates. (b) Free-standing sNBs. Notice the general case of non-homogeneous distribution of the dissolved gas. By dividing the bubble into slices of thickness (dz) and by ignoring the gas concentration in the liquid, the bottom slices show the largest contribution to gas exchange according to Equation (12).
Figure 5. (a) Stabilization of an NB pinned on a surface. As the height and the contact angle decrease, the radius of the curvature increases and the pressure across the interface deflates. (b) Free-standing sNBs. Notice the general case of non-homogeneous distribution of the dissolved gas. By dividing the bubble into slices of thickness (dz) and by ignoring the gas concentration in the liquid, the bottom slices show the largest contribution to gas exchange according to Equation (12).
Nanomaterials 11 02592 g005
Figure 6. The electrical double layer for bNBs; Ψδ is the potential at the boundary between the compact and diffuse layers.
Figure 6. The electrical double layer for bNBs; Ψδ is the potential at the boundary between the compact and diffuse layers.
Nanomaterials 11 02592 g006
Figure 7. Buoyancy rise (cyan line) versus Brownian diffusion (purple line) for bubbles of different sizes. Cross-over point at R = 500 nm and υ = 1 μm/s.
Figure 7. Buoyancy rise (cyan line) versus Brownian diffusion (purple line) for bubbles of different sizes. Cross-over point at R = 500 nm and υ = 1 μm/s.
Nanomaterials 11 02592 g007
Figure 8. (a) A schematic representation of a bulk NB displacement, alternating from an oversaturated domain (blue) to an undersaturated domain (yellow) at equal times but with opposite perturbation conditions. Solid circles reflect the original size of the bubble, and broken circles indicate either expansion or contraction. Notice that the size of the average bubble is invariant. (b) The bubble performs a Brownian walk of an alternating Peano curve of fractal dimension H = 2.
Figure 8. (a) A schematic representation of a bulk NB displacement, alternating from an oversaturated domain (blue) to an undersaturated domain (yellow) at equal times but with opposite perturbation conditions. Solid circles reflect the original size of the bubble, and broken circles indicate either expansion or contraction. Notice that the size of the average bubble is invariant. (b) The bubble performs a Brownian walk of an alternating Peano curve of fractal dimension H = 2.
Nanomaterials 11 02592 g008
Table 1. Dissolution and expansion of different size bubbles at various concentrations.
Table 1. Dissolution and expansion of different size bubbles at various concentrations.
Roc < csTime
to R = 0
c > csTime
to R = 10 Ro
10 μm0.75 cs5 s1.25 cs495 s
1 μm100 ms5 s
100 nm500 μs50 ms
RoccsTime
to R = 0
c ≈ csTime
to R = 10 Ro
10 μm0.9999cs3.5 h1.000114 days
1 μm125 s3.4 h
100 nm1.25 s124 s
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kyzas, G.Z.; Mitropoulos, A.C. From Bubbles to Nanobubbles. Nanomaterials 2021, 11, 2592. https://doi.org/10.3390/nano11102592

AMA Style

Kyzas GZ, Mitropoulos AC. From Bubbles to Nanobubbles. Nanomaterials. 2021; 11(10):2592. https://doi.org/10.3390/nano11102592

Chicago/Turabian Style

Kyzas, George Z., and Athanasios C. Mitropoulos. 2021. "From Bubbles to Nanobubbles" Nanomaterials 11, no. 10: 2592. https://doi.org/10.3390/nano11102592

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop