ArticlePDF Available

Aqueous Nitrogen-Nanobubble dispersion and supersaturation at elevated pressures up to 277 bara

Authors:
Aqueous Nitrogen-Nanobubble Dispersion and Supersaturation at Elevated 1
Pressures up to 277 Bara 2
Tesleem Lawal, Hao Wang, Ryosuke Okuno*, The University of Texas at Austin 3
*Corresponding author: Ryosuke Okuno, Ph.D. 4
Center for Subsurface Energy and the Environment, The University of Texas at Austin 5
200 E. Dean Keeton Street, Stop C0300, Austin, Texas 78712, U.S.A. 6
Phone: +1-512-471-3250; E-mail: okuno@utexas.edu 7
ABSTRACT 8
Aqueous nanobubble (NB) dispersion of gaseous species has been studied and applied for 9
various processes near atmospheric pressure, but its fundamental properties are not well 10
understood at elevated pressures. This paper presents an experimental program that measures the 11
gas content of aqueous NB dispersion of nitrogen (N2) at pressures up to 277 bara. The parameters 12
directly set in the experiments were temperature, external pressure, and total volume while the 13
overall composition of aqueous NB dispersion was obtained by constant mass expansion with 14
material balance. The experimental data were analyzed by using a thermodynamic model for 15
calculating an internally consistent set of properties at the experimental conditions. 16
Results show that the N2 content in aqueous NB dispersion increased significantly with 17
the system pressure. For example, the N2 content at 277 bara was 0.29 mol/L for aqueous NB 18
dispersion with deionized (DI) water, which was 2.9 times greater than the inherent solubility of 19
N2 in DI water at the same pressure. The effect of salinity was studied by using 0.88 M NaCl brine 20
in place of DI water, but the N2 contents were similar to those with DI water for the pressures 21
tested in this research. Application of a thermodynamic model using the GERG-2008 equation of 22
state to the experimental data indicates that NBs themselves were unlikely the main storage of N2, 23
but the existence of NBs enabled the supersaturation of the aqueous phase by N2 because of 24
capillary pressure. 25
26
Keywords: Nanobubble; Supersaturation; Nitrogen; Solubility; Capillary pressure; Phase 27
equilibrium 28
29
1. Introduction 30
Aqueous nanobubble (NB) dispersion has been studied to increase the amount of the water-31
immiscible gaseous species (e.g., nitrogen, oxygen, and carbon dioxide) in an aqueous fluid 32
beyond its inherent solubility. Two types of aqueous NB dispersions are surface nanobubbles and 33
bulk nanobubbles. While the former refers to gas-filled cavities attached to surfaces in the form of 34
spherical caps, the latter refers to gas-filled cavities freely suspended within a bulk liquid [1,2]. 35
This paper is concerned with bulk NBs in the context of aqueous nanobubble dispersions. 36
NBs are classified as having diameters smaller than 1 μm. Studies have shown that their 37
size depends on various parameters, such as gaseous species, pressure, pH, and other operating 38
conditions of their generation [3–5]. The unique properties of NBs, such as enhanced gas 39
concentration beyond the inherent solubility, kinetic stability, and enhanced interfacial area, have 40
made them the subject of research in various fields [6–10]. Most industrial applications are for 41
low-pressure open systems, such as agricultural applications [11], wastewater treatments [12–16], 42
surface cleaning, and separation of materials [17–20]. Comprehensive reviews of NB generation 43
techniques exist in the literature [3, 10, 21–24]. Commonly used techniques include cavitation 44
(acoustic and hydrodynamic) [25–27], water electrolysis [28–30], compression–decompression 45
[31–33], gas injection through porous (glass, ceramic, stainless steel) membranes [34–36], 46
repeated compression of microbubbles [37,38], and mixing of gas and water [39]. 47
The existence and long-term stability of NBs in low-pressure open systems have been 48
discussed in the literature; for example, Alheshibri et al. [1] introduced the concept of Laplace 49
Pressure Bubble Catastrophe (LBPC) regarding unstable nanobubbles. They suggested that, based 50
on classical theories like the Young-Laplace equation and Epstein-Plesset theory [40], the 51
significant pressure difference between the inside and outside of nanobubbles would cause them 52
to dissolve rapidly in milliseconds or less. As the bubble radius decreases, the Laplace pressure 53
increases exponentially, leading to accelerated dissolution. 54
Long-term stability of NBs has been observed even for low-pressure open systems for days 55
[7,41,42], weeks [25,43,44], and months [26,36,45]. Light scattering techniques (e.g., dynamic 56
light scattering or DLS and nanoparticle tracking analysis or NTA) [20,26,32,36,37,41], resonant 57
mass measurements [46,47], high-resolution imaging techniques (e.g., transmission electron 58
microscopy and scanning electron microscopy) [5,48,49], and spectral techniques have been used 59
to verify the existence of NBs [25,39]. These techniques, however, have been limited in 60
differentiating between nanobubbles and nanoparticles. They are also not easily adapted to 61
measurements of high-pressure aqueous NB samples because of the requirement of specially 62
designed experiment rigs. 63
Most discussions on NB stability were made for open systems near atmospheric pressure, 64
and many did not even specify the system for their analysis. Unlike those studied in the literature, 65
the current paper is focused on a high-pressure closed system of aqueous NB dispersion. As will 66
be shown in this paper, the enhanced gas content by aqueous NB dispersion tends to be more 67
significant at a higher pressure (therefore, a closed system); hence, this research was motivated by 68
potential applications of aqueous NB dispersion for subsurface processes, such as enhanced oil 69
recovery (EOR) and geological carbon sequestration (GCS). For example, carbonated water 70
injection has been studied for EOR [50–62] and also has attracted attention for CO2 in-situ 71
mineralization as GCS since the CarbFix project in Iceland [63–67]. An obvious limitation of 72
carbonated water is the CO2 concentration limited by the inherent solubility (e.g., 1 mol/L). 73
Although aqueous NB dispersion has the potential to enhance the gas content in aqueous 74
fluid as widely studied for open systems near atmospheric pressure, its fundamental properties are 75
not well understood especially at elevated pressures relevant to subsurface applications. This is at 76
least in part because it is not easy to directly measure properties of aqueous NB dispersion under 77
elevated pressure, such as gas content, bubble size, bubble number density, phase composition, 78
capillary pressure, and interfacial tension, which are expected to vary for different pressure, 79
temperature, gas species, salinity, and operation conditions of NB generation. Hence, there is a 80
critical need to generate fundamental experimental data on aqueous NB dispersion at elevated 81
pressures. To this end, we started by using nitrogen (N2), instead of CO2, as the gaseous species 82
in this research since CO2 dissolution in water gives additional complexities associated with a pH 83
change. 84
This paper presents an experimental program that measures the gas content of aqueous N2 85
NB dispersion at pressures up to 277 bara and analyzes the data using a thermodynamic model. 86
The parameters directly set in the experiments were temperature, external pressure, and total 87
volume, while the overall composition of aqueous NB dispersion was obtained by constant mass 88
expansion. The limited amount of data that were directly measured was supplemented by a 89
thermodynamic model, which generated a consistent set of properties, including those not directly 90
measurable, such as bubble size, number density, and capillary pressure. The depressurized 91
samples were subjected to measurement of bubble size distribution and bubble number density 92
using DLS and NTA. 93
To the best of our knowledge, this is the first experimental study of NB generation and gas 94
content measurement at high pressures, addressing a critical research gap in the literature. 95
Furthermore, the enhancement of gas content in water in the form of nanobubble dispersions holds 96
immense potential for accelerating storage security in GCS applications and enhancing oil 97
recovery. 98
99
2. Experimental methods 100
2.1. Materials 101
The aqueous samples used in the experiment were deionized (DI) water and sodium 102
chloride (NaCl) brine with a salinity of 50,000 ppm (equivalent to 0.88 M NaCl). The DI water 103
had a resistivity of 18.2 MΩ-cm. NaCl salt (Thermo Scientific Chemicals, purity of 99.0%) was 104
dissolved in DI water to formulate the brine solution. The aqueous solutions were checked for 105
impurities using a NanoSight NS500 instrument and no particles (impurities) were confirmed. 106
Nitrogen gas of research-grade (Linde Gas & Equipment, purity of 99.9999%) was used. 107
2.2. Gas content measurement 108
The N2 gas content measurement was performed in two stages. The first stage was the 109
generation and preparation of the NB sample while the second stage was the depressurization and 110
measurement of the N2 gas content. Several variables – pressure, salinity, volumetric co-injection 111
ratio, total injection flow rate, and membrane pore diameter were tested for the gas content 112
measurements. Table 1 shows the variables investigated for the gas content measurements. All the 113
experiments were conducted at room temperature (295.15 K). The control experiments used DI 114
water, a co-injection ratio of 50% N2 + 50% DI water, a total injection flow rate of 100 mL/hr, a 115
membrane pore diameter of 5 μm, and pressures ranging from 35 to 277 bara. 116
117
Table 1. Experiment variables for the gas content measurement. 118
Variable Variations Other conditions
Salinity DI water, NaCl brine 35 – 277 bara; 5 μm
Co-injection ratio
50% N2; 50% DI water (or
NaCl brine) 35 – 277 bara; 5 μm
60% N2; 40% DI water 70, 139, 208 bara; 5 μm
Total injection flow rate 100 mL/hr 35 – 277 bara; 5 μm
500 mL/hr 70, 139, 208 bara; 5 μm
Membrane pore diameter 2 μm 70, 139, 208 bara
119
Stage 1: Generation and preparation of NB sample 120
Figure 1 shows a schematic of the experimental setup for generating the N2 NB. The setup 121
consists of five accumulators: two accumulators for N2 and DI water (or NaCl brine) and three 122
receiver accumulators; three stainless-steel porous membranes; one Hassler-type core holder with 123
a 70 Durometer Viton sleeve to house the membranes; spacers on either side of the membranes; a 124
hydraulic manual pump to maintain overburden pressure on the core holder; pressure gauges to 125
monitor experiment pressures; Teledyne ISCO syringe pumps to control pressure and flow rates; 126
and a sapphire visualization cell. 127
128
129
Figure 1. Schematic of the experimental setup used to generate the aqueous N2 NB fluid. 130
131
The stainless-steel porous membrane had a porosity of 37%, an average pore size of 5 μm 132
(with a maximum size of 10 μm), an outer diameter of 25.4 mm, and a length of 3 mm. The spacers 133
on either side of the membranes had fluid dispersion channels engraved on the cross-sectional 134
surface to allow for full distribution of fluid across the entire face of the membranes. A confining 135
pressure of 35 bara greater than the experiment pressure was applied to the core holder for each 136
experiment. The pressure gauges were zeroed before each experiment to ensure accurate and 137
consistent pressure readings across all experiments. The sapphire cell was used to visually monitor 138
the behavior of the NB sample and has a known volume (including all connection lines), Vcell, of 139
13.69 mL. It can withstand pressures up to 350 bara and temperatures up to 423.15 K. To prevent 140
gas trapping in any of its connections, the sapphire cell was configured to simulate a single-inlet 141
single-outlet cell by connecting valves to all its connection ports. Upon completion of each 142
experiment, the entire system was thoroughly cleaned with DI water and dried with air before 143
proceeding to the next experiment. 144
Before generating the NB, the entire system was evacuated for 1 hour. The system was 145
saturated with DI water (or NaCl brine) up to receiver accumulator #1 (2c in Figure 1) at the 146
experiment pressure. The sapphire cell and the top of receiver accumulators #2 and #3 (2d and 2e 147
in Figure 1) were filled with N2 at the experiment pressure. To generate the NB, N2 and DI water 148
(or NaCl brine) were co-injected (per the different configurations shown in Table 1) at a constant 149
flow rate through the porous membranes at the specified pressure, co-injection ratio, and total 150
injection flow rate. The co-injected fluids were received downstream of the porous membranes at 151
a constant refill flow rate (maintaining the system pressure) into receiver accumulator #1 until a 152
steady state was reached. After reaching a steady state (approximately 3 minutes for 500 mL/hr 153
and 15 minutes for 100 mL/hr), the co-injected fluids were then received by receiver accumulator 154
#2 at a constant refill flow rate. The duration of the co-injection was 60 minutes for 100 mL/hr and 155
12 minutes for 500 mL/hr. We posit that the formation of nanobubbles occurs in two main stages: 156
first, hydrodynamic mixing and gas snap-off across the membrane leads to the creation of 157
dispersed N2 bubbles in the aqueous phase that is supersaturated by N2, and secondly, the bulk 158
microbubbles are compressed to form bulk nanobubbles under elevated pressure conditions. 159
After the co-injection period, receiver accumulator #2 contained excess bulk N2 as well as 160
the N2 NB. The contents of receiver #2 were then injected through the sapphire cell into receiver 161
#3 at a constant flow rate. First, the N2 gas initially in the sapphire cell was displaced by the excess 162
bulk N2 from receiver #2. Then, the bulk N2 was further displaced by the N2 NB. We ensured the 163
cell contained only N2 NB with no trapped gas by injecting at least twice the cell volume of N2 164
NB into the cell. After filling the cell with the N2 NB, the top and bottom valves of the cell were 165
closed to isolate the cell from the other parts of the system. The sapphire cell then contained the 166
aqueous NB fluid sample at pressure, P1. 167
Stage 2: Depressurization and gas content measurement 168
Figure 2 shows a schematic of the experimental setup for measuring the gas content in the 169
generated N2 NB. The setup consists of a sapphire cell containing the NB sample, a piston 170
accumulator with a pressure gauge, and a glass beaker. 171
172
173
Figure 2. Schematic of the experimental setup to measure the gas content in the aqueous N2 NB 174
fluid. 175
176
The sapphire cell was connected to the accumulator via a stainless-steel tubing. The piston 177
of the accumulator was set at the top of the accumulator. The dead volumes of the tubing and the 178
accumulator were known. Before depressurizing the N2 NB sample, the tubing and dead volume 179
of the accumulator were evacuated for 15 minutes. Then, the N2 NB sample was gradually 180
depressurized. First, the top valve of the sapphire cell was slowly opened to fill the tubing with 181
depressurized N2 escaping the NB sample. Then, the top valve of the accumulator was slowly 182
opened to allow further escape of N2 from the NB sample into the accumulator. Finally, the bottom 183
valve of the accumulator was opened to gauge the displacement of the piston as the aqueous NB 184
fluid sample was depressurized or expanded. The water collected corresponded to the volume of 185
depressurized N2 in the aqueous NB fluid. The mass of the water collected is denoted as mw2, and 186
the pressure at the top side of the accumulator after depressurization is P2. Finally, the remaining 187
water (or brine) in the cell was collected and its mass mw3 was measured. The atmospheric 188
pressure, P3, during the experiment was 1.0135 bara. 189
The N2 gas content was calculated on the following basis: 190
1. There was a negligible amount of N2 in the remaining water phase in the sapphire cell because 191
the solubility of N2 in water at atmospheric pressure is small and because the presence of any 192
bubbles has a negligible impact on the gas content as shown later for NTA data. 193
2. The amount of water was corrected for the transfer of water to the expanded gas phase as mist. 194
The gas content, represented in units of mole per liter (mole of N2 in a volume of aqueous NB 195
fluid), was then calculated as shown below: 196
𝑥(𝑚𝑜𝑙/𝐿)=mole of 𝑁 (mol)
volume of NB sample (L) (1) 197
mole of N2(mol) =𝜌𝑉
𝑀𝑊=󰇧 𝜌
𝑀𝑊󰇨𝑃𝑉𝑍
𝑃𝑍(2) 198
where ρ1 is the density of N2 at the experimental pressure, V1 is the volume of N2 in the aqueous 199
NB fluid at experimental pressure, MW is the molecular weight of N2, P1 is the experimental 200
pressure, P2 is the depressurized pressure, V2 is the total volume of N2 at atmospheric pressure, Z1 201
is the compressibility factor of N2 at the experimental pressure, and Z2 is the compressibility factor 202
of N2 at atmospheric pressure. 203
The gas content data was compared to the thermodynamic solubility of N2 in water at 204
298.15 K [68] and N2 in NaCl brine at 303.15 K [69,70] to determine the solubility enhancement 205
factor of aqueous NB fluids. Studies showed that N2 solubility in NaCl brine at low molarities 206
(e.g., 1 M) is insensitive to temperature [69–71]. For each configuration shown in Table 1, the 207
procedure remained the same. The repeatability of the experiment was tested by repeating the 208
experiments three times at 104 bara for DI water. The standard deviation from these experiments 209
was applied to the remaining experiments as error bars. 210
2.3. Bubble size measurement for depressurized samples 211
The size distribution, number density, mean, and mode of the N2 NB were measured for 212
depressurized samples at atmospheric pressure using two light scattering techniques: dynamic light 213
scattering (DLS) and nanoparticle tracking analysis (NTA). The samples for the bubble size 214
measurement were generated using the same procedure described in Section 2.2. Upon isolating 215
the sapphire cell, however, the sample in the cell was depressurized to atmospheric pressure 216
similarly to the method of compression–decompression [31–33] for low-pressure bulk nanobubble 217
generation. The depressurized sample containing N2 NBs at atmospheric pressure was transferred 218
into glass vials and stored at room temperature. Bubble size measurements were performed 1, 2, 219
7, and 10 days after the NB formation. On day 1, the measurements were performed 1 hour after 220
depressurization, and on the subsequent days, the measurements were performed 26, 147, and 217 221
hours after depressurization. Before measuring the size distribution of the samples, DLS and NTA 222
measurements of the DI water sample were performed three times, and no particles were detected. 223
The DLS measurements used a ZEN3500 Zetasizer Nano ZS particle size analyzer 224
(Malvern Panalytical Ltd.). DLS is a non-invasive scattering technique that measures particle sizes 225
by capturing the random change in intensity of light scattered by particles undergoing Brownian 226
motion within an aqueous medium. The fluctuations in scattered light intensity are then translated 227
into particle sizes using the Stokes-Einstein equation: 228
𝑑=𝑘𝑇
3πη𝐷 (3) 229
where dh is the hydrodynamic diameter, kB is the Boltzmann constant, T is the temperature, η is 230
the dynamic viscosity of the aqueous phase, and D is the diffusion coefficient of particles in the 231
aqueous phase. 232
The Zetasizer Nano ZS can determine a particle size distribution within the range from 4 233
nm to 3 μm. The experiments were performed at 295.15 K to maintain consistent conditions and 234
with a refractive index of 1.33 for water and 1.0 for N2. The particle intensity distribution is 235
determined directly from the measurements and is then mathematically transformed into a number 236
distribution. The measurements were taken at a fixed scattering angle of 173 degrees and repeated 237
twice for each sample. 238
The NTA measurements used a NanoSight NS500 (Malvern Panalytical Ltd.). The NTA 239
technique, built upon dark-field microscopy (DFM), tracks individual particle trajectories to 240
compute sizes based on the Stokes-Einstein equation. The measurements were performed at room 241
temperature and the settings for screen gain, focus, camera level, and detection threshold were 242
optimized to ensure accurate detection of particles while minimizing noise. Each sample was 243
recorded three times and the duration of each capture was 30 seconds. One advantage of NTA over 244
DLS is its capability to determine the number density of particles in an aqueous medium. 245
The DLS and NTA measurements in this research were only for depressurized samples of 246
aqueous NB dispersion, and their application to high-pressure samples is currently not available. 247
Although the measurement conditions are different from the envisioned applications to high-248
pressure processes, the DLS and NTA data are useful to validate the samples of aqueous NB 249
dispersion and to observe the transient behavior of depressurized samples in an open system. Also, 250
low-pressure data serve as a reference for high-pressure data with a correlative capability of a 251
thermodynamic model in this research. 252
253
3. Results and discussion 254
3.1. Effect of pressure and salinity 255
Tables 2 and 3 show the measured gas content data in the experiments for aqueous N2 NB 256
fluid in DI water and 50,000 ppm NaCl brine, respectively. Figure 3 presents the gas content data 257
in mol/L for DI water and NaCl brine at a constant co-injection ratio of 50% N2 + 50% DI water 258
(or NaCl brine), a total injection flow rate of 100 mL/hr, and a 5 μm membrane. The red squares 259
and blue triangles in Figure 3 represent the inherent solubility of N2 in water and NaCl brine, 260
respectively. The error bars in Figure 3 represent an estimate of the variability in the experimental 261
measurements. Note that xN2 in Tables 2 and 3 and Figure 3 represents the overall N2 concentration 262
including the dispersion of N2 bubbles and the dispersion of N2 molecules in the external aqueous 263
phase. 264
Table 2. Measured gas content data for aqueous N2 NB in DI water at 295.15 K. 265
P1
(bara)
P2
(bara)
mw2
(g)
mw3
(g)
Vcell
(mL)
xN2
Inherent
solubility (mol/L)
Solubility
enhancement
34.59
1.22
4.54
10.765 ± 3
13.69
0.0515
0.0112
4.59
68.93
1.29
11.4
10.930 ± 3
13.69
0.0801
0.0226
3.54
104.2
1.15
23.58
11.405 ± 3
13.69
0.1118
0.0432
2.59
103.3
1.29
19.49
11.310 ± 3
13.69
0.1098
0.0425
2.58
103.4
1.29
21.22
11.295 ± 3
13.69
0.1165
0.0425
2.74
138.2
1.22
32.01
11.040 ± 3
13.69
0.1505
0.0526
2.86
207.8
1.22
57.41
11.834 ± 3
13.69
0.2401
0.0796
3.02
277.1
1.22
71.92
12.040 ± 3
13.69
0.2922
0.1000
2.92
266
Table 3. Measured gas content data for aqueous N2 NB in 50,000 ppm NaCl brine at 295.15 K. 267
P1
(bara)
P2
(bara)
mw2
(g)
mb3
(g)
Vcell
(mL)
xN2
Inherent
solubility (mol/L)
Solubility
enhancement
47.07
1.22
3.53
11.155 ± 3
13.69
0.0463
0.0167
2.77
69.82
1.22
9.01
11.850 ± 3
13.69
0.0637
0.0258
2.47
103.0
1.22
19.98
10.800 ± 3
13.69
0.1075
0.0255
4.22
138.5
1.29
27.62
11.530 ± 3
13.69
0.1402
0.0349
4.02
207.8
1.29
43.66
11.800 ± 3
13.69
0.2008
0.0509
3.94
274.0
1.36
66.55
11.980 ± 3
13.69
0.3037
0.0648
4.69
268
The results show that as pressure increased from 35 to 277 bara, the gas content increased 269
significantly for both DI water and NaCl brine. Note that this figure shows the N2 concentration 270
on a logarithmic scale. The gas content in aqueous NB dispersion with NaCl brine was close to 271
that with DI water across all the pressures tested. Thus, the 0.88 M NaCl brine had no significant 272
impact on the N2 gas content in the aqueous NB sample. The N2 concentration in the N2 NB 273
dispersion with NaCl brine was 2.5 times greater than the inherent solubility at 69.8 bara, and the 274
enhancement factor was as large as 4.7 at 274.0 bara as shown in Table 3. An extrapolation of the 275
gas content data to atmospheric pressure for DI water gives an N2 gas content of 0.0121 mol/L (or 276
4.7 × 10-4 mole fraction of N2). 277
278
279
Figure 3. N2 gas content in DI water and 50,000 ppm NaCl brine. The red squares represent the 280
thermodynamic solubility of N2 in water [68]. The blue triangles represent the thermodynamic 281
solubility of N2 in NaCl brine [69,70]. 282
283
3.2. Effect of coinjection ratio, total injection rate, and membrane pore diameter 284
Figure 4 shows the effects of co-injection ratio, total injection flow rate, and membrane 285
pore diameter at different pressures on the gas content of N2 in the aqueous NB sample (DI water). 286
Overall, the results show negligible impacts of co-injection ratio, total injection flow rate, and 287
membrane pore diameter on the N2 gas content compared to the baseline experiments. These 288
results are in contradiction to the results reported for low-pressure experiments [4,35,45] about the 289
effect of changing the co-injection ratios or pore membrane diameters on gas content data. This is 290
likely because these effects were significant and noticeable at low pressures and then became less 291
significant at high pressures; that is, increasing the pressure had the most significant impact on the 292
N2 gas content in DI water and NaCl brine among the experimental parameters tested in this 293
research. 294
295
296
Figure 4. N2 gas content in DI water at different operating conditions (co-injection ratio, total 297
injection flow rate, and membrane pore diameter) and different pressures. The baseline case used 298
a co-injection rate of 50% N2 + 50% DI water through a 5 μm membrane at a total injection flow 299
rate of 100 mL/hr. 300
301
3.3. Bubble size measurement using DLS and NTA 302
Figure 5 shows the size distributions of depressurized samples for aqueous NB dispersion 303
of N2 in DI water based on DLS. Table 4 shows the mean and mode bubble sizes, polydispersity 304
indices (PDI), and diffusion coefficients from the DLS measurements on days 1, 2, 7, and 10. 305
Before the measurements, the DI water used to generate the NB was checked for the presence of 306
nanosized impurities and none were detected. 307
308
Table 4. Mean, mode, PDI, and diffusion coefficients of the aqueous N2 NB sample using DLS. 309
Day Bubble size,
mean (nm)
Bubble size,
mode (nm)
PDI Diffusion
coefficient (m
2
/s)
1
235.1 ±
7.4
3
158.6
±
5.45
0.354 ±
0.05
(
1.93
±
0.0
6
)
×
10
-12
2 169.6 ± 5.09 132.4 ± 14.9 0.112 ± 0.16
(
2.67
±
0.08
)
×
10
-12
7
212.3 ±
6.08
134.0
±
20.9
0.359 ±
0.01
(
2.13
±
0.06
)
×
10
-12
10 156.7 ± 10.9 107.7 ± 7.28 0.243 ± 0.26
(
2.90
±
0.
20
)
×
10
-12
310
311
Figure 5. Size distribution of the aqueous N2 NB sample using DLS. 312
313
The DLS data show notable trends in bubble size over time. On day 1, the mean bubble 314
size was 235.1 nm, which decreased to 169.6 nm on day 2, followed by an increase to 212.3 nm 315
on day 7, and a subsequent decrease to 156.7 nm on day 10. The mode of the bubble size was 316
158.6, 132.4, 134.0, and 107.7 nm for days 1, 2, 7, and 10, respectively. The difference in behavior 317
between the mean and mode bubble sizes comes from the transient distribution of bubble sizes as 318
given in Figure 5. The PDI data also reflect the transient size distribution using DLS. On day 1, 319
the PDI was 0.354, correlating to a mean size significantly larger than the mode size. This suggests 320
the presence of a substantial number of larger bubbles contributing to the scattered light signal. 321
However, on day 2, the PDI dropped to 0.112, indicating the absence of significantly larger 322
bubbles. On day 7, the PDI increased to 0.359, possibly indicating the coalescence of smaller 323
bubbles into larger bubbles, causing the mean size to become greater than the mode size. Finally, 324
on day 10, the PDI decreased again to 0.243, further confirming the reduction in the number of 325
larger particles, and thus a reduction in the mean bubble size. The PDI behavior gives credence to 326
the presence of gas-filled nanobubbles which tend to coalesce into larger bubbles and eventually 327
dissipate either through bursting or rising to the surface. 328
Figure 6 shows the laser-illuminated N2 nanobubbles using NTA at atmospheric pressure 329
on days 1, 2, 7, and 10. The detected nanobubbles are shown as white dots on a black background. 330
One advantage of NTA over DLS is its capability to determine a bubble number density in the 331
sample. Figure 7 shows the size distribution of the N2 NB in DI water. Table 5 gives the mean and 332
mode bubble sizes, bubble number densities, and diffusion coefficients from the NTA 333
measurements. The transient behavior of NB size distribution is also shown in Figure 8. Like the 334
DLS measurements, the DI water used to generate the NB for the NTA measurements was checked 335
for the presence of nanosized impurities and no impurities were detected. 336
Figures 7 and 8 show that the overall behavior of aqueous NB dispersion became more 337
stabilized over time, suggesting that the numerical results given in Tables 4 and 5 should be 338
carefully understood together with the overall trends shown in Figures 5, 7, and 8. 339
Table 5. Mean, mode, bubble number density, and diffusion coefficients of the aqueous N2 NB 340
sample using NTA. 341
Day Bubble size,
mean (nm)
Bubble size,
mode (nm)
Bubble number density
(bubbles/mL)
Diffusion
coefficient (m
2
/s)
1
101.8 ± 11.9
97.9
0
± 14.6
(9.25 ± 0.63)
×
10
7
(
4.74
± 0.
92
)
×
10
-12
2 116.5 ± 20.4 103.2 ± 11.0
(2.56 ± 1.39)
×
10
8
(
4.20
±
1.12
)
×
10
-12
7
103.2 ± 6.30
84.7
0
± 12.5
(5.85 ± 2.18)
×
10
7
(
4.45
±
0.47
)
×
10
-12
10 128.0 ± 17.8 102.1 ± 8.60
(1.72 ± 0.69)
×
10
7
(3.67 ± 0.78)
×
10
-12
342
(a)
Day 1
(b)
Day 2
(c)
Day 7
(d)
Day 10
Figure 6. Laser-illuminated N2 NB in DI water using NTA. 343
344
345
Figure 7. Size distribution of the aqueous N2 NB sample using NTA. 346
347
348
Figure 8. NB number density of the aqueous N2 NB samples using NTA. 349
350
On day 1, the mean bubble size was 101.8 nm, and the mode bubble size was 97.9 nm with 351
a bubble concentration of 9.25 × 107 bubbles/mL. There was a close agreement between the mean 352
and mode bubble sizes. On day 2, both the mean and mode bubble sizes increased to 116.5 nm and 353
103.2 nm, respectively, with a notably higher number density of 2.56 × 108 bubbles/mL. The 354
results from days 1 and 2 show a significant level of transient behavior of aqueous NB dispersion 355
with the possibilities of bubble coalescence, growth, and nucleation over time. On day 7, the mean 356
bubble size slightly decreased to 103.2 nm, and the mode bubble size decreased to 84.7 nm. The 357
number density was 5.85 × 107 bubbles/mL. Finally, on day 10, the mean bubble size increased to 358
128.0 nm, while the mode bubble size also increased to 102.1 nm with a number density of 1.72 × 359
107 bubbles/mL. 360
Figure 9 compares DLS and NTA measurements at atmospheric pressure in terms of mean 361
and mode bubble sizes. NTA measurements show a closer agreement between the mean and mode 362
sizes than DLS measurements. This is because of the inherent difference in the measurement 363
principles of both measurement techniques. That is, NTA measures the rate of Brownian motion 364
of the bubbles while DLS measures the fluctuations in the scattered light intensity of the bubbles 365
undergoing Brownian motion. This fundamental difference leads to NTA being less susceptible to 366
polydispersity, thus, showing a closer agreement between the mean and mode bubble sizes. 367
368
369
Figure 9. Mean and mode of the aqueous N2 NB sample using DLS and NTA. 370
371
3.4. Thermodynamic analysis of aqueous NB dispersion of N2 372
The most important benefit of using thermodynamics to analyze the experimental data in 373
this research is that a thermodynamic model can give an internally consistent set of properties for 374
the entire experimental conditions, which enables understanding the overall behavior of aqueous 375
NB dispersion. Also, such properties include unmeasurable ones in high-pressure experiments, 376
such as phase composition and capillary pressure. As described previously, the experimental 377
procedure directly specifies the temperature, total volume, and pressure of a closed system of 378
aqueous nanobubble dispersion of N2. Then, the constant mass expansion of the system gives the 379
mole numbers of water and N2. With these experimentally measurable variables, the most 380
convenient approach to analyzing the experimental data is to minimize the Helmholtz free energy 381
using a thermodynamic model [72]. Then, a solution for the minimization problem gives 382
thermodynamic properties for the dispersion of N2-rich nanobubbles in the aqueous phase that is 383
supersaturated by N2 at the experimental conditions. 384
The traditional two-phase system with a planar interface requires NC + 2 variables to be 385
set, where NC is the number of components; however, the current problem requires setting NC + 3 386
variables. The additional variable comes from the generalization to allow for different pressures 387
for the two phases (i.e., capillary pressure). In this research, the NC + 3 variables are temperature, 388
total volume, aqueous-phase pressure, and mole numbers of the components (water and N2). With 389
these specifications, this section describes properties of the aqueous NB dispersions generated in 390
the experiments using a single thermodynamic model, which gives internal consistency of the 391
calculated properties. We recognize limitations of the modeling; for example, the binary system 392
(water and N2) requires a uniform size of bubbles, and the mole numbers of water and N2 (from 393
the constant mass expansion) contain uncertainty. Therefore, thermodynamic properties given for 394
aqueous NB dispersions in this section must be understood as apparent values satisfying 395
experimentally specified conditions and material balance. 396
The modeling approach largely follows Achour and Okuno [72] and Achour et al. [73], but 397
we have extended it to cover all experimental conditions in this research. The GERG-2008 398
equation of state (EOS, [74]) was used because the EOS used in this research should be reliable in 399
constructing the Helmholtz free energy surface in metastable regions [75,76]; i.e., the aqueous 400
phase supersaturated by N2. The EOS was first calibrated using experimental data for water-N2 at 401
298.15, 308.15, 318.15, and 323.15 K from 80 to 1015 bara [77]. Then, it was calibrated for NaCl 402
brine-N2 using Smith et al. [69] and O’Sullivan and Smith [70], in which 1 M NaCl brine was used 403
at 303.15, 324.65, 375.65, and 398.15 K from 12 to 608 bara. Although the experimental 404
temperature in this research, 295.15 K, is not contained in the data of Smith et al. [69] and 405
O’Sullivan and Smith [70], the N2 solubility in NaCl brine at small molarities (e.g., 1 M) is 406
insensitive to temperature. Also, the NaCl concentration in this research, 0.88 M, is close to the 407
NaCl concentration, 1 M, for the data. Figure 10 compares the calculated results using the GERG-408
2008 EOS with the experimental data in the literature for water-N2 and NaCl brine-N2. Results 409
show good agreement between the EOS results and the data; therefore, the GERG-2008 EOS was 410
successfully calibrated. Supplementary Material gives a detailed description of the GERG-2008 411
EOS model and parameters for the water-N2 and NaCl brine-N2 mixtures. 412
413
414
Figure 10. Calibrated GERG-2008 EOS model for the N2/water [77,78] and N2/NaCl brine 415
[69,70] solubility data at different pressures and temperatures. 416
417
The experimental data given in Tables 2 and 3 were then used to solve the calibrated EOS 418
for properties of aqueous NB dispersion of N2 for the given temperature, total volume, mole 419
numbers for water/brine and N2, and external-phase pressure. Among these input variables, the 420
water mole numbers based on mw3 (Table 2) and mb3 (Table 3) were quite influential to the 421
resulting diameter of bubbles. Tables 2 and 3 show that the uncertainty of mw3 (and mb3) was ±3 422
g. Even a minor variation (e.g., 0.1 g) in mw3 (or mb3) significantly influenced the calculated 423
apparent bubble diameter by one order of magnitude; therefore, it was not possible to quantitatively 424
determine an apparent bubble diameter for these data with an order-of-magnitude accuracy using 425
thermodynamic calculations only. Therefore, we allowed for adjustment in mole numbers of the 426
two components within the uncertainty range such that the Helmholtz free energy was minimized 427
while satisfying the specified temperature, total volume, and external pressure. We confirmed that 428
such adjustments resulted in overall compositions within the uncertainty range stated above. 429
Figures 11 and 12 show the results from the GERG-2008 EOS, such as bubble diameter, 430
bubble number density, supersaturation of aqueous phase by N2, the fraction of N2 as bubble 431
dispersion, and capillary pressure, for the N2 NB fluid in DI water and NaCl brine, respectively. 432
The error bars were determined as half the difference between the smallest and largest values at 433
104 bara (for DI water) since the experiment was repeated three times at this pressure. For the 434
aqueous supersaturation levels (Figures 11c and 12c), the N2 concentration in the aqueous phase 435
was compared with the inherent solubility of N2 based on the GERG-2008 EOS. For Figures 11d 436
and 12d, the mole number of N2 as bubbles was divided by the total mole number of N2 for each 437
sample. 438
Figures 11ab and 12ab show that the calculated diameter decreased, and the number density 439
of bubbles increased with increasing external pressure. For the DI water case, the extrapolation to 440
atmospheric pressure using the two lowest pressure data yields a bubble diameter of 158 nm. This 441
extrapolated diameter lies within the range of measured bubble diameters, 50-400 nm, at 442
atmospheric pressure in the literature [1,21]. Also see Figures 5, 7, and 9 for the data measured in 443
this research. Figures 11cd and 12cd show that a large fraction of the N2 in the system is 444
molecularly dissolved in the external aqueous phase. That is, the existence of bubbles increases 445
the N2 content in the system by increasing the molecule dispersion (supersaturation) in the aqueous 446
phase much more than by containing N2 as bubbles. The presence of bubbles in aqueous NB fluid 447
tends to increase the level of supersaturation in the external aqueous phase with capillary pressure 448
(Figures 11e and 12e). The results indicate that this supersaturation is the main contribution to the 449
amount of N2 in the aqueous NB fluid, which has fundamental impacts on the research and 450
development of NB technologies. 451
Figure 3 shows that the N2 content did not substantially change between DI water and NaCl 452
brine. However, comparison of Figures 11 and 12 indicates that the NaCl salinity may have 453
reduced the amount of N2 that was molecularly dissolved, which required increasing the amount 454
of N2 as bubbles to satisfy the material balance in the closed system of the experiments in this 455
research. 456
457
Figure 11. Bubble diameter, bubble number density, N2 mole fraction, fraction of N2 contained in 458
bubbles, and capillary pressure of aqueous NB fluid in DI water at 295.15 K for pressures up to 459
277 bara. 460
461
462
Figure 12. Bubble diameter, bubble number density, N2 mole fraction, fraction of N2 contained in 463
bubbles, and capillary pressure of aqueous NB fluid in NaCl brine at 295.15 K for pressures up 464
to 274 bara. 465
466
467
4. Conclusions 468
This paper presented a new set of experimental data for aqueous NB dispersions of N2 at elevated 469
pressures of up to 277 bara using N2 and DI water or 0.88 M NaCl brine with specially designed 470
stainless-steel porous membranes. The N2 content in the generated NB sample was investigated 471
for different parameters, such as pressure, salinity, co-injection ratio, total injection flow rate, and 472
membrane pore diameter. The experimental results were supplemented with the GERG-2008 EOS 473
model to give a qualitative overview of the characteristics of the aqueous NB fluid. The main 474
conclusions are as follows: 475
1. The N2 content in aqueous NB dispersion increased significantly with pressure. At 277 bara, 476
for example, the inherent solubility of N2 in DI water is 0.10 mol/L, but the N2 content was 477
enhanced to 0.29 mol/L by aqueous NB dispersion with DI water. With 0.88 M NaCl brine, 478
the inherent solubility of N2 is 0.065 mol/L at 274 bara, but the N2 content was enhanced to 479
0.30 mol/L by aqueous NB dispersion. The impact of NaCl salinity on the N2 content in 480
aqueous NB dispersion was not observed for high-pressure experiments in this research. 481
2. The N2 content in aqueous NB dispersion was insensitive to injection parameters, such as the 482
total injection rate and the co-injection ratio, in this research. In particular, using a 2 μm 483
membrane resulted in a similar gas content as using a 5 μm membrane. 484
3. Analysis of the experimental data using the GERG-2008 EOS showed that the apparent 485
diameter of N2 NBs ranged from 14 to 120 nm for DI water and 5 to 115 nm for 0.88 M NaCl 486
brine under the experimental conditions. Additionally, the analysis indicated that NBs were 487
not the main storage of N2, but they enabled the supersaturation of the aqueous phase by N2 488
enhancing the N2 content in the system. 489
4. The DLS and NTA data measured for depressurized samples of aqueous NB dispersion of N2 490
in this research showed transient behavior over 10 days. Number densities and size 491
distributions of NBs were consistent with the reported data in the literature. 492
493
Declaration of Competing Interest 494
The authors declare that they have no known competing financial interests or personal 495
relationships that could have appeared to influence the work reported in this paper. 496
497
Acknowledgements 498
We gratefully acknowledge the members of the Energi Simulation Industrial Affiliate Program on 499
Carbon Utilization and Storage (ES Carbon UT) at the Center for Subsurface Energy and the 500
Environment at the University of Texas at Austin for their support. Ryosuke Okuno holds the 501
Pioneer Corporation Faculty Fellowship in Petroleum Engineering at the University of Texas at 502
Austin. 503
504
Nomenclature 505
Abbreviations 506
CO2 carbon dioxide 507
DFM dark-field microscopy 508
DI deionized 509
DLS dynamic light scattering 510
EOR enhanced oil recovery 511
EOS equation of state 512
GCS geological carbon sequestration 513
LPBC Laplace Pressure Bubble Catastrophe 514
N2 nitrogen 515
NaCl sodium chloride 516
NB nanobubble 517
NC number of components 518
NTA nanoparticle tracking analysis 519
PDI polydispersity index 520
521
Symbols 522
D diffusion coefficient 523
dh hydrodynamic diameter 524
kB Boltzmann constant 525
m mass 526
MW molecular weight 527
P pressure 528
T temperature 529
V volume 530
x gas content 531
Z compressibility factor 532
ρ density 533
η dynamic viscosity 534
535
Subscripts 536
b brine 537
cell sapphire cell 538
w water 539
540
REFERENCES 541
1. Alheshibri, M.; Qian, J.; Jehannin, M.; Craig, V. S. J. A History of Nanobubbles. Langmuir 542
2016, 32 (43), 11086–11100. https://doi.org/10.1021/acs.langmuir.6b02489. 543
2. Tan, B. H.; An, H.; Ohl, C.-D. Stability of Surface and Bulk Nanobubbles. Current Opinion in 544
Colloid & Interface Science 2021, 53, 101428. https://doi.org/10.1016/j.cocis.2021.101428. 545
3. Temesgen, T.; Bui, T. T.; Han, M.; Kim, T.; Park, H. Micro and Nanobubble Technologies as 546
a New Horizon for Water-Treatment Techniques: A Review. Advances in Colloid and 547
Interface Science 2017, 246, 40–51. https://doi.org/10.1016/j.cis.2017.06.011. 548
4. Meegoda, J. N.; Aluthgun Hewage, S.; Batagoda, J. H. Stability of Nanobubbles. 549
Environmental Engineering Science 2018, 35 (11), 1216–1227. 550
https://doi.org/10.1089/ees.2018.0203. 551
5. Jadhav, A. J.; Barigou, M. Bulk Nanobubbles or Not Nanobubbles: That Is the Question. 552
Langmuir 2020, 36 (7), 1699–1708. https://doi.org/10.1021/acs.langmuir.9b03532. 553
6. Ebina, K.; Shi, K.; Hirao, M.; Hashimoto, J.; Kawato, Y.; Kaneshiro, S.; Morimoto, T.; 554
Koizumi, K.; Yoshikawa, H. Oxygen and Air Nanobubble Water Solution Promote the Growth 555
of Plants, Fishes, and Mice. PLoS ONE 2013, 8 (6), e65339. 556
https://doi.org/10.1371/journal.pone.0065339. 557
7. Oh, S. H.; Han, J. G.; Kim, J.-M. Long-Term Stability of Hydrogen Nanobubble Fuel. Fuel 558
2015, 158, 399–404. https://doi.org/10.1016/j.fuel.2015.05.072. 559
8. Perez Sirkin, Y. A.; Gadea, E. D.; Scherlis, D. A.; Molinero, V. Mechanisms of Nucleation 560
and Stationary States of Electrochemically Generated Nanobubbles. J. Am. Chem. Soc. 2019, 561
141 (27), 10801–10811. https://doi.org/10.1021/jacs.9b04479. 562
9. Batchelor, D. V. B.; Armistead, F. J.; Ingram, N.; Peyman, S. A.; Mclaughlan, J. R.; Coletta, 563
P. L.; Evans, S. D. Nanobubbles for Therapeutic Delivery: Production, Stability and Current 564
Prospects. Current Opinion in Colloid & Interface Science 2021, 54, 101456. 565
https://doi.org/10.1016/j.cocis.2021.101456. 566
10. Favvas, E. P.; Kyzas, G. Z.; Efthimiadou, E. K.; Mitropoulos, A. C. Bulk 567
Nanobubbles, Generation Methods and Potential Applications. Current Opinion in Colloid & 568
Interface Science 2021, 54, 101455. https://doi.org/10.1016/j.cocis.2021.101455. 569
11. Pal, P.; Joshi, A.; Anantharaman, H. Nanobubble Ozonation for Waterbody Rejuvenation at 570
Different Locations in India: A Holistic and Sustainable Approach. Results in Engineering 571
2022, 16, 100725. https://doi.org/10.1016/j.rineng.2022.100725. 572
12. Oturan, M. A.; Aaron, J.-J. Advanced Oxidation Processes in Water/Wastewater Treatment: 573
Principles and Applications. A Review. Critical Reviews in Environmental Science and 574
Technology 2014, 44 (23), 2577–2641. https://doi.org/10.1080/10643389.2013.829765. 575
13. Atkinson, A. J.; Apul, O. G.; Schneider, O.; Garcia-Segura, S.; Westerhoff, P. Nanobubble 576
Technologies Offer Opportunities To Improve Water Treatment. Acc. Chem. Res. 2019, 52 577
(5), 1196–1205. https://doi.org/10.1021/acs.accounts.8b00606. 578
14. Lyu, T.; Wu, S.; Mortimer, R. J. G.; Pan, G. Nanobubble Technology in Environmental 579
Engineering: Revolutionization Potential and Challenges. Environ. Sci. Technol. 2019, 53 580
(13), 7175–7176. https://doi.org/10.1021/acs.est.9b02821. 581
15. Levitsky, I.; Tavor, D.; Gitis, V. Micro and Nanobubbles in Water and Wastewater Treatment: 582
A State-of-the-Art Review. Journal of Water Process Engineering 2022, 47, 102688. 583
https://doi.org/10.1016/j.jwpe.2022.102688. 584
16. Jia, M.; Farid, M. U.; Kharraz, J. A.; Kumar, N. M.; Chopra, S. S.; Jang, A.; Chew, J.; Khanal, 585
S. K.; Chen, G.; An, A. K. Nanobubbles in Water and Wastewater Treatment Systems: Small 586
Bubbles Making Big Difference. Water Research 2023, 245, 120613. 587
https://doi.org/10.1016/j.watres.2023.120613. 588
17. Liu, G.; Wu, Z.; Craig, V. S. J. Cleaning of Protein-Coated Surfaces Using Nanobubbles: An 589
Investigation Using a Quartz Crystal Microbalance. J. Phys. Chem. C 2008, 112 (43), 16748–590
16753. https://doi.org/10.1021/jp805143c. 591
18. Liu, G.; Craig, V. S. J. Improved Cleaning of Hydrophilic Protein-Coated Surfaces Using the 592
Combination of Nanobubbles and SDS. ACS Appl. Mater. Interfaces 2009, 1 (2), 481–487. 593
https://doi.org/10.1021/am800150p. 594
19. Zhang, M.; Seddon, J. R. T. Nanobubble–Nanoparticle Interactions in Bulk Solutions. 595
Langmuir 2016, 32 (43), 11280–11286. https://doi.org/10.1021/acs.langmuir.6b02419. 596
20. Zhu, J.; An, H.; Alheshibri, M.; Liu, L.; Terpstra, P. M. J.; Liu, G.; Craig, V. S. J. Cleaning 597
with Bulk Nanobubbles. Langmuir 2016, 32 (43), 11203–11211. 598
https://doi.org/10.1021/acs.langmuir.6b01004. 599
21. Zhou, L.; Wang, S.; Zhang, L.; Hu, J. Generation and Stability of Bulk Nanobubbles: A Review 600
and Perspective. Current Opinion in Colloid & Interface Science 2021, 53, 101439. 601
https://doi.org/10.1016/j.cocis.2021.101439. 602
22. Han, G.; Chen, S.; Su, S.; Huang, Y.; Liu, B.; Sun, H. A Review and Perspective on Micro and 603
Nanobubbles: What They Are and Why They Matter. Minerals Engineering 2022, 189, 604
107906. https://doi.org/10.1016/j.mineng.2022.107906. 605
23. Foudas, A. W.; Kosheleva, R. I.; Favvas, E. P.; Kostoglou, M.; Mitropoulos, A. C.; Kyzas, G. 606
Z. Fundamentals and Applications of Nanobubbles: A Review. Chemical Engineering 607
Research and Design 2023, 189, 64–86. https://doi.org/10.1016/j.cherd.2022.11.013. 608
24. Li, X.; Peng, B.; Liu, Q.; Liu, J.; Shang, L. Micro and Nanobubbles Technologies as a New 609
Horizon for CO2-EOR and CO2 Geological Storage Techniques: A Review. Fuel 2023, 341, 610
127661. https://doi.org/10.1016/j.fuel.2023.127661. 611
25. Ohgaki, K.; Khanh, N. Q.; Joden, Y.; Tsuji, A.; Nakagawa, T. Physicochemical Approach to 612
Nanobubble Solutions. Chemical Engineering Science 2010, 65 (3), 1296–1300. 613
https://doi.org/10.1016/j.ces.2009.10.003. 614
26. Nirmalkar, N.; Pacek, A. W.; Barigou, M. On the Existence and Stability of Bulk Nanobubbles. 615
Langmuir 2018, 34 (37), 10964–10973. https://doi.org/10.1021/acs.langmuir.8b01163. 616
27. Oliveira, H.; Azevedo, A.; Rubio, J. Nanobubbles Generation in a High-Rate Hydrodynamic 617
Cavitation Tube. Minerals Engineering 2018, 116, 32–34. 618
https://doi.org/10.1016/j.mineng.2017.10.020. 619
28. Kikuchi, K.; Tanaka, Y.; Saihara, Y.; Maeda, M.; Kawamura, M.; Ogumi, Z. Concentration of 620
Hydrogen Nanobubbles in Electrolyzed Water. Journal of Colloid and Interface Science 2006, 621
298 (2), 914–919. https://doi.org/10.1016/j.jcis.2006.01.010. 622
29. Kikuchi, K.; Nagata, S.; Tanaka, Y.; Saihara, Y.; Ogumi, Z. Characteristics of Hydrogen 623
Nanobubbles in Solutions Obtained with Water Electrolysis. Journal of Electroanalytical 624
Chemistry 2007, 600 (2), 303–310. https://doi.org/10.1016/j.jelechem.2006.10.005. 625
30. Kikuchi, K.; Ioka, A.; Oku, T.; Tanaka, Y.; Saihara, Y.; Ogumi, Z. Concentration 626
Determination of Oxygen Nanobubbles in Electrolyzed Water. Journal of Colloid and Interface 627
Science 2009, 329 (2), 306–309. https://doi.org/10.1016/j.jcis.2008.10.009. 628
31. Fang, Z.; Wang, L.; Wang, X.; Zhou, L.; Wang, S.; Zou, Z.; Tai, R.; Zhang, L.; Hu, J. 629
Formation and Stability of Surface/Bulk Nanobubbles Produced by Decompression at Lower 630
Gas Concentration. J. Phys. Chem. C 2018, 122 (39), 22418–22423. 631
https://doi.org/10.1021/acs.jpcc.8b05688. 632
32. Ke, S.; Xiao, W.; Quan, N.; Dong, Y.; Zhang, L.; Hu, J. Formation and Stability of Bulk 633
Nanobubbles in Different Solutions. Langmuir 2019, 35 (15), 5250–5256. 634
https://doi.org/10.1021/acs.langmuir.9b00144. 635
33. Xu, W.; Wang, Y.; Huang, Q.; Wang, X.; Zhou, L.; Wang, X.; Wen, B.; Guan, N.; Hu, J.; 636
Zhou, X.; Zhang, L. The Generation and Stability of Bulk Nanobubbles by Compression-637
Decompression Method: The Role of Dissolved Gas. Colloids and Surfaces A: 638
Physicochemical and Engineering Aspects 2023, 657, 130488. 639
https://doi.org/10.1016/j.colsurfa.2022.130488. 640
34. Kukizaki, M.; Goto, M. Size Control of Nanobubbles Generated from Shirasu-Porous-Glass 641
(SPG) Membranes. Journal of Membrane Science 2006, 281 (1–2), 386–396. 642
https://doi.org/10.1016/j.memsci.2006.04.007. 643
35. Ahmed, A. K. A.; Sun, C.; Hua, L.; Zhang, Z.; Zhang, Y.; Zhang, W.; Marhaba, T. Generation 644
of Nanobubbles by Ceramic Membrane Filters: The Dependence of Bubble Size and Zeta 645
Potential on Surface Coating, Pore Size and Injected Gas Pressure. Chemosphere 2018, 203, 646
327–335. https://doi.org/10.1016/j.chemosphere.2018.03.157. 647
36. Ulatowski, K.; Sobieszuk, P.; Mróz, A.; Ciach, T. Stability of Nanobubbles Generated in Water 648
Using Porous Membrane System. Chemical Engineering and Processing - Process 649
Intensification 2019, 136, 62–71. https://doi.org/10.1016/j.cep.2018.12.010. 650
37. Jin, J.; Feng, Z.; Yang, F.; Gu, N. Bulk Nanobubbles Fabricated by Repeated Compression of 651
Microbubbles. Langmuir 2019, 35 (12), 4238–4245. 652
https://doi.org/10.1021/acs.langmuir.8b04314. 653
38. Jin, J.; Wang, R.; Tang, J.; Yang, L.; Feng, Z.; Xu, C.; Yang, F.; Gu, N. Dynamic Tracking of 654
Bulk Nanobubbles from Microbubbles Shrinkage to Collapse. Colloids and Surfaces A: 655
Physicochemical and Engineering Aspects 2020, 589, 124430. 656
https://doi.org/10.1016/j.colsurfa.2020.124430. 657
39. Oh, S. H.; Kim, J.-M. Generation and Stability of Bulk Nanobubbles. Langmuir 2017, 33 (15), 658
3818–3823. https://doi.org/10.1021/acs.langmuir.7b00510. 659
40. Epstein, P. S.; Plesset, M. S. On the Stability of Gas Bubbles in Liquid‐Gas Solutions. 1950. 660
41. Wang, Q.; Zhao, H.; Qi, N.; Qin, Y.; Zhang, X.; Li, Y. Generation and Stability of Size-661
Adjustable Bulk Nanobubbles Based on Periodic Pressure Change. Sci Rep 2019, 9 (1), 1118. 662
https://doi.org/10.1038/s41598-018-38066-5. 663
42. Hewage, S. A.; Kewalramani, J.; Meegoda, J. N. Stability of Nanobubbles in Different Salts 664
Solutions. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2021, 609, 665
125669. https://doi.org/10.1016/j.colsurfa.2020.125669. 666
43. Ushikubo, F. Y.; Furukawa, T.; Nakagawa, R.; Enari, M.; Makino, Y.; Kawagoe, Y.; Shiina, 667
T.; Oshita, S. Evidence of the Existence and the Stability of Nano-Bubbles in Water. Colloids 668
and Surfaces A: Physicochemical and Engineering Aspects 2010, 361 (1–3), 31–37. 669
https://doi.org/10.1016/j.colsurfa.2010.03.005. 670
44. Azevedo, A.; Etchepare, R.; Calgaroto, S.; Rubio, J. Aqueous Dispersions of Nanobubbles: 671
Generation, Properties and Features. Minerals Engineering 2016, 94, 29–37. 672
https://doi.org/10.1016/j.mineng.2016.05.001. 673
45. Alam, H. S.; Sutikno, P.; Soelaiman, T. A. F.; Sugiarto, A. T. Bulk Nanobubbles: Generation 674
Using a Two-Chamber Swirling Flow Nozzle and Long-Term Stability in Water. J Flow Chem 675
2022, 12 (2), 161–173. https://doi.org/10.1007/s41981-021-00208-8. 676
46. Alheshibri, M.; Craig, V. S. J. Armoured Nanobubbles; Ultrasound Contrast Agents under 677
Pressure. Journal of Colloid and Interface Science 2019, 537, 123–131. 678
https://doi.org/10.1016/j.jcis.2018.10.108. 679
47. Alheshibri, M.; Jehannin, M.; Coleman, V. A.; Craig, V. S. J. Does Gas Supersaturation by a 680
Chemical Reaction Produce Bulk Nanobubbles? Journal of Colloid and Interface Science 681
2019, 554, 388–395. https://doi.org/10.1016/j.jcis.2019.07.016. 682
48. Uchida, T.; Oshita, S.; Ohmori, M.; Tsuno, T.; Soejima, K.; Shinozaki, S.; Take, Y.; Mitsuda, 683
K. Transmission Electron Microscopic Observations of Nanobubbles and Their Capture of 684
Impurities in Wastewater. Nanoscale Res Lett 2011, 6 (1), 295. https://doi.org/10.1186/1556-685
276X-6-295. 686
49. Rabinowitz, J.; Whittier, E.; Liu, Z.; Jayant, K.; Frank, J.; Shepard, K. Nanobubble-Controlled 687
Nanofluidic Transport. Science Advances 2020, 6 (46), eabd0126. 688
https://doi.org/10.1126/sciadv.abd0126. 689
50. Christensen, R. J. Carbonated Waterflood Results--Texas And Oklahoma. Annual Meeting of 690
Rocky Mountain Petroleum Engineers of AIME, Farmington, New Mexico, USA, 1961; SPE-691
66-MS. https://doi.org/10.2118/66-MS. 692
51. Blackford, T. A. Carbonated Waterflood Implementation and Its Impact on Material 693
Performance in a Pilot Project. SPE Annual Technical Conference and Exhibition, Dallas, 694
Texas, 1987; SPE-16831-MS. https://doi.org/10.2118/16831-MS. 695
52. Perez, J. M. Carbonated Water Imbibition Flooding: An Enhanced Oil Recovery Process for 696
Fractured Reservoirs. 1992, 12. 697
53. Dong, Y.; Dindoruk, B.; Ishizawa, C.; Lewis, E.; Kubicek, T. An Experimental Investigation 698
of Carbonated Water Flooding. SPE Annual Technical Conference and Exhibition, Denver, 699
Colorado, USA, 2011; SPE-145380-MS. https://doi.org/10.2118/145380-MS. 700
54. Riazi, M.; Sohrabi, M.; Jamiolahmady, M. Experimental Study of Pore-Scale Mechanisms of 701
Carbonated Water Injection. Transp Porous Med 2011, 86 (1), 73–86. 702
https://doi.org/10.1007/s11242-010-9606-8. 703
55. Sohrabi, M.; Kechut, N. I.; Riazi, M.; Jamiolahmady, M.; Ireland, S.; Robertson, G. Safe 704
Storage of CO2 Together with Improved Oil Recovery by CO2-Enriched Water Injection. 705
Chemical Engineering Research and Design 2011, 89 (9), 1865–1872. 706
https://doi.org/10.1016/j.cherd.2011.01.027. 707
56. Sohrabi, M.; Emadi, A.; Farzaneh, S. A.; Ireland, S. A Thorough Investigation of Mechanisms 708
of Enhanced Oil Recovery by Carbonated Water Injection. SPE Annual Technical Conference 709
and Exhibition, Houston, Texas, USA, 2015; SPE-175159-MS. 710
https://doi.org/10.2118/175159-MS. 711
57. Foroozesh, J.; Jamiolahmady, M.; Sohrabi, M. Mathematical Modeling of Carbonated Water 712
Injection for EOR and CO2 Storage with a Focus on Mass Transfer Kinetics. Fuel 2016, 174, 713
325–332. https://doi.org/10.1016/j.fuel.2016.02.009. 714
58. Seyyedi, M.; Mahzari, P.; Sohrabi, M. A Comparative Study of Oil Compositional Variations 715
during CO2 and Carbonated Water Injection Scenarios for EOR. Journal of Petroleum Science 716
and Engineering 2018, 164, 685–695. https://doi.org/10.1016/j.petrol.2018.01.029. 717
59. Esene, C.; Rezaei, N.; Aborig, A.; Zendehboudi, S. Comprehensive Review of Carbonated 718
Water Injection for Enhanced Oil Recovery. Fuel 2019, 237, 1086–1107. 719
https://doi.org/10.1016/j.fuel.2018.08.106. 720
60. Bisweswar, G.; Al-Hamairi, A.; Jin, S. Carbonated Water Injection: An Efficient EOR 721
Approach. A Review of Fundamentals and Prospects. J Petrol Explor Prod Technol 2020, 10 722
(2), 673–685. https://doi.org/10.1007/s13202-019-0738-2. 723
61. Ajoma, E.; Sungkachart, T.; Saira, -; Yin, H.; Le-Hussain, F. A Laboratory Study of 724
Coinjection of Water and CO2 to Improve Oil Recovery and CO2 Storage: Effect of Fraction 725
of CO2 Injected. SPE Journal 2021, 26 (04), 2139–2147. https://doi.org/10.2118/204464-PA. 726
62. Talebi, A.; Hasan-Zadeh, A.; Kazemzadeh, Y.; Riazi, M. A Review on the Application of 727
Carbonated Water Injection for EOR Purposes: Opportunities and Challenges. Journal of 728
Petroleum Science and Engineering 2022, 214, 110481. 729
https://doi.org/10.1016/j.petrol.2022.110481. 730
63. Liu, D.; Agarwal, R.; Li, Y.; Yang, S. Reactive Transport Modeling of Mineral Carbonation 731
in Unaltered and Altered Basalts during CO2 Sequestration. International Journal of 732
Greenhouse Gas Control 2019, 85, 109–120. https://doi.org/10.1016/j.ijggc.2019.04.006. 733
64. Snæbjörnsdóttir, S. Ó.; Sigfússon, B.; Marieni, C.; Goldberg, D.; Gislason, S. R.; Oelkers, E. 734
H. Carbon Dioxide Storage through Mineral Carbonation. Nat Rev Earth Environ 2020, 1 (2), 735
90–102. https://doi.org/10.1038/s43017-019-0011-8. 736
65. Marieni, C.; Voigt, M.; Clark, D. E.; Gíslason, S. R.; Oelkers, E. H. Mineralization Potential 737
of Water-Dissolved CO2 and H2S Injected into Basalts as Function of Temperature: Freshwater 738
versus Seawater. International Journal of Greenhouse Gas Control 2021, 109, 103357. 739
https://doi.org/10.1016/j.ijggc.2021.103357. 740
66. Raza, A.; Glatz, G.; Gholami, R.; Mahmoud, M.; Alafnan, S. Carbon Mineralization and 741
Geological Storage of CO2 in Basalt: Mechanisms and Technical Challenges. Earth-Science 742
Reviews 2022, 229, 104036. https://doi.org/10.1016/j.earscirev.2022.104036. 743
67. Wang, H.; Carrasco-Jaim, O.; Okuno, R. Aqueous Nanobubble Dispersion of CO2 in Sodium 744
Formate Solution for Enhanced CO2 Mineralization Using Basaltic Rocks, the 2024 Carbon 745
Capture, Utilization, and Storage conference, March 11 – 13, 2024, Houston, Texas. 746
68. Wiebe, R.; Gaddy, V. L.; Heins, C. Solubility of Nitrogen in Water at 25°C from 25 to 1000 747
Atmospheres. Ind. Eng. Chem. 1932, 24 (8), 927–927. https://doi.org/10.1021/ie50272a023. 748
69. Smith, N. O.; Kelemen, S.; Nagy, B. Solubility of Natural Gases in Aqueous Salt Solutions—749
II. Geochimica et Cosmochimica Acta 1962, 26 (9), 921–926. https://doi.org/10.1016/0016-750
7037(62)90066-2. 751
70. O’Sullivan, T. D.; Smith, N. O. Solubility and Partial Molar Volume of Nitrogen and Methane 752
in Water and in Aqueous Sodium Chloride from 50 to 125.Deg. and 100 to 600 Atm. J. Phys. 753
Chem. 1970, 74 (7), 1460–1466. https://doi.org/10.1021/j100702a012. 754
71. O’Sullivan, T. D.; Smith, N. O.; Nagy, B. Solubility of Natural Gases in Aqueous Salt 755
Solutions—III Nitrogen in Aqueous NaCl at High Pressures. Geochimica et Cosmochimica 756
Acta 1966, 30 (6), 617–619. https://doi.org/10.1016/0016-7037(66)90015-9. 757
72. Achour, S. H.; Okuno, R. Phase Stability Analysis for Tight Porous Media by Minimization of 758
the Helmholtz Free Energy. Fluid Phase Equilibria 2020, 520, 112648. 759
https://doi.org/10.1016/j.fluid.2020.112648. 760
73. Achour, S. H.; Sheng, K.; Lawal, T.; Okuno, R. Thermodynamic Modeling of Aqueous 761
Nanobubble Dispersion. SPE Annual Technical Conference and Exhibition, San Antonio, 762
Texas, USA, 2023. https://doi.org/10.2118/215122-MS. 763
74. Kunz, O.; Wagner, W. The GERG-2008 Wide-Range Equation of State for Natural Gases and 764
Other Mixtures: An Expansion of GERG-2004. J. Chem. Eng. Data 2012, 57 (11), 3032–3091. 765
https://doi.org/10.1021/je300655b. 766
75. Imre, A. R.; Baranyai, A.; Deiters, U. K.; Kiss, P. T.; Kraska, T.; Quiñones Cisneros, S. E. 767
Estimation of the Thermodynamic Limit of Overheating for Bulk Water from Interfacial 768
Properties. Int J Thermophys 2013, 34 (11), 2053–2064. https://doi.org/10.1007/s10765-013-769
1518-8. 770
76. Aursand, P.; Gjennestad, M. Aa.; Aursand, E.; Hammer, M.; Wilhelmsen, Ø. The Spinodal of 771
Single- and Multi-Component Fluids and Its Role in the Development of Modern Equations of 772
State. Fluid Phase Equilibria 2017, 436, 98–112. https://doi.org/10.1016/j.fluid.2016.12.018. 773
77. Maslennikova, V. Y. Solubility of Nitrogen in Water. Tr. Gos. Nauchno-Issled. Proektn. Inst. 774
Azotn. Prom-sti. Prod. Org. Sint. 1971, 12, 82–87. 775
78. Liu, Y.; Hou, M.; Ning, H.; Yang, D; Yang, G; Han, B. Phase Equilibria of CO2 + N2 + H2O 776
and N2 + CO2 + H2O + NaCl + KCl + CaCl2 Systems at Different Temperatures and Pressures. 777
Journal of Chemical & Engineering Data 2012, 57(7), 1928–1932. 778
https://doi.org/10.1021/je3000958. 779
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
The amount of gaseous species in water or brine can be greatly enhanced in the form of nanobubble (NB) dispersion. Aqueous NB dispersion has vast industrial applications, potentially in enhanced oil recovery (EOR) and carbon dioxide (CO2) sequestration to control the mobility of gaseous species and the geochemistry associated with CO2 dissolution. Development of such NB technologies depends on a proper understanding of thermodynamic properties of aqueous NB dispersion. The objectives of this research are to analyze the thermodynamic stability of aqueous NB dispersion and to apply a thermodynamic equilibrium model to analyze experimental data. We first present a thermodynamic formulation for modeling aqueous NB dispersion, which clarifies that aqueous NB dispersion occurs in the aqueous phase that is supersaturated by the gaseous species in the system. That is, the gaseous species are present in two modes—dispersion of gas bubbles under capillary pressure and molecule dispersion (supersaturation) in the external aqueous phase. Such a thermodynamic system is referred to as aqueous NB fluid in this research and specified by (NC + 3) variables (e.g., temperature, total volume, components’ mole numbers, and capillary pressure), in which NC is the number of components. We then present a novel implementation of the GERG-2008 equation of state (EOS) in minimization of the Helmholtz free energy to solve for equilibrium properties of aqueous NB fluid. GERG-2008 was used in this research because it is suitable for modeling an aqueous phase that is supersaturated by gaseous species. The thermodynamic equilibrium model was applied to experimental data of aqueous NB fluid with nitrogen (N2) at pressures up to 277 bara (4,019 psia) and 295.15 K (71.6°F). Application of the model to experimental data indicates that a large fraction (0.8–0.9) of the total amount of N2 is in the form of molecule dispersion, but such supersaturation of the aqueous phase is possible because of the presence of NB dispersion with capillary pressure. That is, NB dispersion can increase the gas content in aqueous NB fluid by enabling gas supersaturation in the aqueous phase as a thermodynamic system.
Conference Paper
The amount of gaseous species in water or brine can be greatly enhanced in the form of nanobubble (NB) dispersion. Aqueous NB dispersion has vast industrial applications, potentially in enhanced oil recovery and carbon dioxide (CO2) sequestration to control the mobility of gaseous species. Development of such NB technologies depends on a proper understanding of thermodynamic properties of aqueous NB dispersion. The objectives of this research are to analyze the thermodynamic stability of aqueous NB dispersion and to apply a thermodynamic equilibrium model to analyze experimental data. We first present a thermodynamic formulation for modeling aqueous NB dispersion, which clarifies that aqueous NB dispersion occurs in the aqueous phase that is supersaturated by the gaseous species in the system. That is, the gaseous species are present in two modes: dispersion of gas bubbles under capillary pressure, and molecule dispersion (supersaturation) in the external aqueous phase. Such a thermodynamic system is referred to as aqueous NB fluid in this research, and specified by (NC + 3) variables (e.g., temperature, total volume, components’ mole numbers, and capillary pressure), in which NC is the number of components. We then present a novel implementation of the GERG-2008 equation of state (EOS) in minimization of the Helmholtz free energy to solve for equilibrium properties of aqueous NB fluid. GERG-2008 was used in this research because it is suitable for modeling an aqueous phase that is supersaturated by gaseous species. The thermodynamic equilibrium model was applied to experimental data of aqueous NB fluid with nitrogen (N2) at pressures up to 277 bara (4019 psia) and 295.15 K (71.6°F). Application of the model to experimental data indicates that a large fraction (0.8 – 0.9) of the total amount of N2 is in the form of molecule dispersion, but such supersaturation of the aqueous phase is possible because of the presence of NB dispersion with capillary pressure. That is, NB dispersion can increase the gas content in aqueous NB fluid by enabling gas supersaturation in the aqueous phase as a thermodynamic system. Although experimental uncertainties resulted in a possible range of equilibrium properties for aqueous NB fluids at high pressures, the extrapolation of the calculation results to atmospheric pressure yielded a radius and a number density of bubbles within the range of data reported in the literature.
Article
Since the discovery of nanobubbles (NBs) in 1994, NBs have been attracting growing attention for their fascinating properties and have been studied for application in various environmental fields, including water and wastewater treatment. However, despite the intensive research efforts on NBs’ fundamental properties, especially in the past five years, controversies and disagreements in the published literature have hindered their practical implementation. So far, reviews of NB research have mainly focused on NBs’ role in specific treatment processes or general applications, highlighting proof-of-concept and success stories primarily at the laboratory scale. As such, there lacks a rigorous review that authenticates NBs’ potential beyond the bench scale. This review aims to provide a comprehensive and up-to-date analysis of the recent progress in NB research in the field of water and wastewater treatment at different scales, along with identifying and discussing the challenges and prospects of the technology. Herein, we systematically analyze (1) the fundamental properties of NBs and their relevancy to water treatment processes, (2) recent advances in NB applications for various treatment processes beyond the lab scale, including over 20 pilot and full-scale case studies, (3) a preliminary economic consideration of NB-integrated treatment processes (the case of NB-flotation), and (4) existing controversies in NBs research and the outlook for future research. This review is organized with the aim to provide readers with a step-by-step understanding of the subject matter while highlighting key insights as well as knowledge gaps requiring research to advance the use of NBs in the wastewater treatment industry.
Article
Nanoscale gas bubbles are paid more and more attention due to their significant applications in different fields including the environmental remediation, plant and animal growth as well as medical diagnosis, etc. As reported, the local gas saturation plays an important role for the formation of surface nanobubbles (NBs) but is less importance for their stability. As for bulk NBs, few researches focused on the influence of dissolved gas on their generation and stability because it is thought generally that a limited amount of gas could dissolve into water. Herein, we reported for the first time the relationship of dissolved gases (Kr, O2 and N2) and the formation and stability of bulk NBs. Firstly, we developed a compression-decompression method to produce the water with super-high concentration of dissolved gas. About 60 mg/L oxygen dissolved gas was created to promote the formation of bulk NBs. It was showed that high bulk NB concentrations were produced using the compression-decompression method by controlling the loading pressure and time at the same time. The evolution process showed that the concentration of dissolved gas would decrease quickly with the deposited time. However, the concentration of formed bulk NBs did not follow the same way as dissolved gas concentration. It exhibited a complicated change over the time. Typically, first sharp increase to one order higher concentration than at the beginning and then decreased with a fluctuation within 72 hours. More interestingly, the time of this sharp increase in nanobubble concentration depended on the type of gas, the krypton (Kr) gas system took longer time to reach the highest concentration and the oxygen (O2) as well as nitrogen (N2) gas system reached the highest concentration at about 4 h generally. The change of zeta potential of those NBs followed the same fluctuation as their concentration. Finally, we presumed a theoretical model to explain the evolution mechanism of bulk NBs. It indicated that there is a competition of different bubble behaviors (nucleation, clustering and coalescence) in different time periods. This study provides a new technique to produce high concentration of bulk NBs and dissolved gas in solution. Those results are very significance for further understanding the mechanism of formation and stability of bulk NBs under a super-high concentration of dissolved gas and may be used in some chemical reactions related with gas to promote the reaction efficiency.
Article
In this study, four different sites of ponds at different time period were analysed to show to the effect of Nanobubble ozonation (NBO). NBO is one of novel technologies which is environmentally friendly only using air, ozone, and requiring virtually no chemicals. All the case studies have been performed in real time in different seasons and locations which gives the idea of importance of NBO for the treatment of water bodies with completely natural, environmentally sustainable, cost effective, and less time-consuming technology. Three case studies have been investigated in-situ (Site 1, 2, 4) and one ex-situ (site-3), which verify the efficacy of NBO in treating pond water. NBO treated water was determined to be in the standard limits, odour was eliminated, and water was cleaned enough to be consumed by the animals. The test results show that ozone nanobubble (NBO) in wastewater treatments achieved 85–99% reduction in total soluble solids (TSS), 80–90% reduction in biochemical oxygen demand (BOD), and 55% chemical oxygen demand (COD) reduction at site 3 and 82% COD reduction at site 4. Improved dissolved oxygen suitable for the living beings was achieved. Hence, this paper emphasizes the efficacy of NBO treatment to reclaim the water bodies and ecological restoration and to achieve the sustainable goals of clean water and environmental sustainability. DO level in all the ponds improved significantly after NBO treatment and showed the value of 14.5 mg/L when measured even after 50 h. Nanobubble gas dissolution system not only able to improve the dissolved oxygen up to supersaturation level but also able to retain it consistent more than 14–15 h. The vary reason for maintaining such consistency is the size of nanobubbles which is around <0.5 μm and NB doesn't follow the buoyancy, hence move in water in Brownian motion. NBG systems are future of our lakes and ponds rejuvenation and maintain the water quality.
Article
Over the past two decades, there has been a rapid growth in the research and extraordinary applications of micro and nanobubbles due to their unique physicochemical properties beyond conventional bubbles including large specific surface area, slow rising velocity, high interface potential, high mass transfer efficiency, and so on. Although there were considerable excellent results and findings reported, a remarkable lag is still recorded concerning the detailed explanation of their formation principle, generating methods and special physicochemical properties. Simultaneously, limited efforts have been made on collecting and preparing a well-organized, comprehensive report of their potential applications in the areas of flotation science and technology. This review is dedicated to a systematic description on their generating principle, fabricating methods, special physicochemical properties and state-of-the-art characterization technologies based on up-to-date research progresses. Subsequently, the main applications of micro and nanobubbles in typical flotation processes with a special emphasis on mineral & secondary resource processing, environment & water treatment, separation & purification for metals were comprehensively reviewed and thoroughly discussed. Ultimately, the challenges and limitations of the current research field including the opportunities for future work of micro and nanobubbles were prospected. Oily-bubble floating-extraction, as the 3th generation foam extraction technology, were also proposed to enhance mass transfer. In summary, this review may be helpful to improve the overall understanding on Micro and Nanobubbles: What They Are and Why They matter.
Article
Climate change is taking place due to significant emissions of greenhouse gases into the atmosphere. CO2 storage in geological formations is a promising approach that can help to reduce greenhouse gas emissions from large emitters such as the steel and cement industries. However, effective storage in underground formations requires active trapping mechanisms to reduce the likelihood of leakage. Carbon mineralization is a trapping technique that can permanently store CO2 in reactive rocks such as basalt. Although this method has been known for a long time, only two pilot projects in Iceland and the USA practiced CO2 injection into basalts. This could be mainly due to the complexity of the interactions, the rapid carbon mineralization, and the difficulty to estimate the storage capacity in the long term. In this paper, we discuss different mechanisms and technical challenges of CO2 storage in igneous rocks and propose a selection criterion based on laboratory and field-scale experience. It appears that basalt is a suitable rock for rapid carbon mineralization given its worldwide distribution, vesicular texture, and favourable mineral composition, but the lack of effective monitoring techniques and the amount of water required for injection are two major challenges that need to be addressed.
Article
The selection of miscible or immiscible injection methods depends on the desired thermodynamic and operating conditions. However, gas injection alone does not increase reservoir replacement for various reasons. The reason for this is the greater mobility of gas than oil and gravitational separation. These reasons cause the injected gas to not have a favorable effect on the reduction of residual oil and causes early breakthroughs. To solve this problem, the researchers changed the method of gas injection, in which different methods such as continuous water and gas injection, foam gas injection and water and gas combination injection were proposed. The use of a combination of carbon dioxide and produced water is due to the higher solubility of this gas in water than other gases in nature and, the lower miscibility pressure. Also, the availability of this gas and the effective environmental effect of using this gas, has caused the research conducted to dissolve gas in water on the issue of combining carbon dioxide with added water. In this article, we will review the research on carbonated water injection as an enhanced oil recovery; provide challenges, solutions, and determining suitable conditions for injection. Also the number of studies conducted in recent years and important topics are mentioned.