Sunteți pe pagina 1din 44

International Reviews in Physical Chemistry

ISSN: 0144-235X (Print) 1366-591X (Online) Journal homepage: https://www.tandfonline.com/loi/trpc20

Insight into the bubble column evaporator and its


applications

Muhammad Shahid, Chao Fan & Richard Mark Pashley

To cite this article: Muhammad Shahid, Chao Fan & Richard Mark Pashley (2016) Insight into the
bubble column evaporator and its applications, International Reviews in Physical Chemistry, 35:1,
143-185, DOI: 10.1080/0144235X.2016.1147144

To link to this article: https://doi.org/10.1080/0144235X.2016.1147144

Published online: 06 Apr 2016.

Submit your article to this journal

Article views: 180

View Crossmark data

Citing articles: 9 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=trpc20
International Reviews in Physical Chemistry, 2016
Vol. 35, No. 1, 143–185, http://dx.doi.org/10.1080/0144235X.2016.1147144

Insight into the bubble column evaporator and its applications


Muhammad Shahid, Chao Fan and Richard Mark Pashley*

School of Physical, Environmental and Mathematical Sciences, University of New South Wales,
Canberra, Australia
(Received 9 November 2015; final version received 22 January 2016)

This paper presents a review of the bubble column evaporator (BCE) and its many
novel applications. The BCE process offers a continuously produced source of high
air–water interface and consequently provides high overall heat and mass transfer
coefficients. Although the bubbling process itself is both simple to use and apply, our
understanding of the fundamental physical and chemical principles involved is sur-
prisingly limited and there are many issues yet to be explained. Recently the process
has been used to develop new methods for the precise determination of enthalpies of
vaporisation (ΔHvap) of concentrated salt solutions, as an evaporative cooling system,
a sub-boiling thermal desalination unit, for sub-boiling thermal sterilisation, for low
temperature thermal decomposition of different solutes in aqueous solution and for
the inhibition of particle precipitation in supersaturated solutions. These novel appli-
cations can be very useful in many industrial practices, such as desalination, water/
waste water treatment, thermolysis of ammonium bicarbonate (NH4HCO3) for the
regeneration in forward osmosis and refrigeration related industries. The background
theories and models used to explain the BCE process are also reviewed and this fun-
damental knowledge is applied to the design of BCE systems and to explain recently
explored applications, as well as potential improvements. Many other prospective
applications of the BCE process are also reported in this paper.
Keywords: bubble column evaporator; sub-boiling; bubble coalescence inhibition;
thermolysis; desalination; enthalpy of vaporisation; supersaturation; sterilisation;
precipitation

Contents PAGE

1. Introduction to the bubble column evaporator 144

2. Theoretical analysis of the BCE 146


2.1. Bubble formation 146
2.1.1. Basic theories 146
2.1.2. Laplace pressure equation 147
2.2. Bubble coalescence and its inhibition 147
2.2.1. Recent theoretical analysis of coalescence inhibition by
electrolyte solutions 148
2.2.2. Simple experimental study to test the effect of dissolved CO2 148
2.2.3. DLVO interaction forces between bubbles 149
2.2.4. Other factors affecting bubble coalescence or break up 149

*Corresponding author. Email: R.Pashley@adfa.edu.au

© 2016 Informa UK Limited, trading as Taylor & Francis Group


144 M. Shahid et al.

2.3. Vaporisation model 150


2.4. Bubble evaporation layer model 153
2.5. Bubble surface hot layer model 153
2.6. Bubble water vapour equilibration 154
2.7. Bubble rise velocity 155
2.8. Thermal energy balance in the BCE 157
2.9. Comparison of the BCE desalination process with multistage flash
distillation and reverse osmosis 160

3. Applications of the BCE 161


3.1. Precise measurements for the determination of enthalpies of
vaporisation (ΔHvap) 161
3.1.1. Background 161
3.1.2. Determination of ΔHvap values for concentrated aqueous solutions 162
3.2. Evaporative cooling 164
3.2.1. Background 164
3.2.2. BCE evaporative cooling 164
3.3. Seawater desalination 166
3.3.1. Background 166
3.3.2. Bubble column desalination 167
3.3.3. Improved bubble column desalination 168
3.3.4. Comparison and benefits 170
3.4. Water sterilisation 171
3.4.1. Background 171
3.4.2. Bubble column sterilisation 171
3.4.3. Proposed mechanism of BCE sterilisation 173
3.5. Thermolysis of solutes in aqueous solution 174
3.5.1. Background 174
3.5.2. Decomposition of NH4HCO3 solutions 175
3.5.3. Decomposition of K2S2O8 solution using a BCE process 175
3.5.4. Proposed mechanism of BCE solute thermolysis 176
3.6. The controlled growth of precipitates in aqueous salt solutions 176
3.6.1. Background 176
3.6.2. Controlled CaSO4·2H2O precipitation in a BCE and
comparison with simple stirring 178
3.6.3. The application of precipitation inhibition with the BCE 179
3.7. Other potential applications 181

4. Conclusion 181

Disclosure statement 183

References 183

1. Introduction to the bubble column evaporator


The bubble column evaporator (BCE) offers a good illustration of the use of a
gas–liquid interface to drive fundamental processes involving heat and mass transfer.
Bubble columns are devices in which a gas, often dry air, is pumped through a multi-
porous sinter disc to form gas bubbles which are continuously replenished and come in
International Reviews in Physical Chemistry 145

intimate contact with the column solution. Dry gas bubbles in the column solution may
be used simply to mix the liquid phase homogenously to attain uniform temperature
distribution or to saturate dissolved gases into the column solution. Substances can also
be transferred from one phase to the other, for example, when liquid reaction products
are stripped from a gas, where both mass and heat transfer processes can occur
simultaneously [1].
More recently, aqueous BCEs have been used for a range of new applications based
largely on the unexpected effects of many salts and sugars on inhibiting bubble-bubble
coalescence in water, in combination with limited bubble rise rates and rapid water
vapour uptake into the bubbles [2–5].
Heat transfer using shell and tube heat exchangers is the fundamental chemical
engineering process by which heat is transported between two fluids via a material hav-
ing high thermal conductivity (i.e. copper, Cu). Nonetheless, heat transfer may also be
accomplished directly by mixing the solution and the heating fluid (for example, water
and hot air bubbles), which characterises the so-called direct-contact evaporators. Col-
lier, first revealed the concept of heat transfer via bubbles in his patent, published in
1887 [6] and by using this concept, the first commercial plant was installed in the USA
in the early twentieth century [7].
The main advantages of bubble columns, using direct-contact heat transfer, com-
pared to other multiphase reactors are: (a) less maintenance required due to the absence
of moving parts, (b) higher effective interfacial areas and overall mass transfer coeffi-
cients can be achieved, (c) higher heat transfer rates per unit volume of the reactors can
be attained, (d) solids can be handled without any erosion or plugging problems, (e) less
floor space is occupied and bubble column reactors are less costly, (f) slow reactions can
be carried out due to high liquid residence time [8], and (g) the product can be recovered
from the reaction mixture without additional separation operations [1].
The BCE method is a simple method in practice but its theoretical analysis is rather
complex and is yet to be understood in a detailed manner. The phenomenon of bubble
coalescence inhibition was first experimentally discovered by Russian mineral flotation
engineers more than 80 years ago, as the addition of certain amount of salt produced
smaller bubbles, and hence the efficiency of flotation was improved [9]. This occurs as
the bubbles originated though a multi-porous glass sinter plate which prevents bubble
accumulation above a sufficient salt concentration. However, even though there is still
no distinct explanation for bubble coalescence inhibition, it has been well studied and
documented [10–12] and the behaviour has recently been applied to the development

Figure 1. Different applications of the BCE.


146 M. Shahid et al.

of a wide range of useful techniques such as a new method for the precise
determination of enthalpies of vaporisation (ΔHvap) of concentrated salt solutions
[3,13], evaporative cooling [4], a new method for thermal desalination [14,15], a novel
method for sub-boiling, thermal sterilisation [16–18], a novel method for the low
temperature thermal decomposition of different solutes in aqueous solution [2] and a
new approach to aqueous precipitation in a controlled manner [19] (see Figure 1). In
addition to these methods, the thermal design of a bubble column condenser has also
been studied for the production of high quality water as condensate [20–22].

2. Theoretical analysis of the BCE


The theoretical analysis of the BCE presented here is based upon the applied physical
laws and theoretical knowledge of bubble formation, bubble coalescence and its inhibi-
tion, a vaporisation model at the air–water interface, a bubble evaporation layer model, a
model of the heated layer around the surface of the bubbles, bubble water vapour capture,
bubble rising velocity, and the steady state thermal energy balance within the BCE.

2.1. Bubble formation


2.1.1. Basic theories
An early theoretical attempt to explain the bubble formation phenomenon was made by
Davidson and Schuler [23] who assumed that the bubble, at any stage of formation,
ascends with a velocity decided by inertial and viscous forces. Then the bubble was
assumed to detach when its centre had covered a distance equal to the radius of the
final bubble. On the other hand, Khurana and Kumar [24] and others [25–28] assumed
that bubbles form in two stages, i.e. the expansion stage and the detachment stage. Dur-
ing the first stage the bubble expands while its base remains attached to the tip of the
orifice; whereas in the detachment stage the bubble base moves away from the tip.
The above two explanations of bubble formation involve various forces that have to
be taken into account. They are the upwards buoyancy force (FB) and a contact pressure
force (FCP) due to the action of the overpressure inside the bubble (and inside the pore,
see Figure 2) acting on the area of the bubble at the orifice [26]; while the downward
forces are the vertical component of the capillary force (FC) due to the surface tension and
the dynamic force (FD) due to inertial and viscous effects. The first stage ends when the
sum of the upward forces becomes equal to the sum of the downward forces. In the sec-
ond stage the upward forces exceed the sum of the downward forces. As a result, the base
of the expanding bubble moves away from the pore opening [26]. The acceleration of the
bubble in this stage is governed by Newton’s second law of motion [28].
The sum of the vertical forces acting on the bubble which influence its upward
momentum is approximately given by the sum of terms:
X
F ¼ FB þ FCP þ FC þ FD (1)
where the upward forces are positive and downward forces are negative.

2.1.2. Laplace pressure equation


The Laplace pressure is the pressure difference between the inside and the outside of a
bubble. The pressure difference is caused by the surface tension of the curved interface
International Reviews in Physical Chemistry 147

Figure 2. (Colour online) Forces acting on a growing bubble (reproduced from S. Bari and A.
Robinson [26]).

Table 1. Required internal pressures for nucleating bubbles in water at 25 °C.


Bubble radius (nm) 1 2 10 1000
Laplace pressure (atm) 1440 720 144 1.44

between liquid and gas. The Laplace pressure is determined from the Young–Laplace
equation given as:
2c
DP ¼ (2)
r
where, ΔP is the pressure difference between inside and outside of a spherical bubble,
γ is the surface tension of water and r is the radius of the bubble. Using this equation,
a 1 mm radius bubble is found to have negligible extra pressure. Yet when the diameter
is ~3 μm, the bubble has an additional pressure of 1 atm greater inside than outside.
When the bubble is only several hundred nanometres, the pressure inside can be several
atmospheres, as shown in Table 1 [29].

2.2. Bubble coalescence and its inhibition


Bubble coalescence behaviour is vital in water-based cleaning processes, ore-flotation,
food processing, gas–oil separations, absorption and distillation [30]. However, bubble
coalescence inhibition was also found to depend critically on salt type [5,31]. Salt solu-
tions that do inhibit coalescence have recently been used in many applications, as
shown in Figure 1. Foulk attempted to explain the theory of bubble coalescence and
break up in 1929 [32] on the basis that both positively and negatively adsorbed dis-
solved matter cause bubble coalescence inhibition in liquids and by contrast, pure liq-
uids do not prevent coalescence. It is also worth noting that like coalescence inhibition,
coalescence is also still considered complicated, since it involves not only interactions
of bubbles with the surrounding liquid, but also bubbles themselves once they are
brought together by the external flow or by their body forces [33].
Generally, three models have been so far proposed for the coalescence process, that
is the (1) film drainage model [32,34], (2) energy model [35,36] and (3) the critical
approach velocity model [37–39]. In all cases, bubble interaction forces and the
energetics of bubble collisions are the prerequisite of coalescence behaviour [34]. On the
148 M. Shahid et al.

other hand, several attempts to explain bubble coalescence inhibition by the addition of
some salts have been based on: (1) the specific properties of particular salts, such as
their unusual effect on solution viscosity [5,31] (2) the generation of a local electro-
static field [11,40] (3) the use of attractive van der Waals forces between bubbles [41]
and (4) long-range hydrophobic forces [41–43]. These forces will determine whether
the film between two colliding bubbles can drain fast enough to allow coalescence
before they are forced apart in the turbulence of the bubble column and it is generally
believed that bubbles will coalesce once the intervening film reaches about 50 nm [40].
The coalescence process is clearly dependent on dynamic effects since bubbles slowly
forced together always coalesce in aqueous salt solutions. The consideration of interac-
tion forces between approaching bubbles should therefore be focussed on determining
the rate of approach and thinning of the intervening water film.

2.2.1. Recent theoretical analysis of coalescence inhibition by electrolyte solutions


Katsir and Marmur [44,45] recently attempted to explain this mystifying phenomenon
on the basis of bubble surface charging, originating from the speciation of dissolved
CO2, which sustains an electric double layer repulsion. They argued that rapid coales-
cence in electrolyte solutions occurred via two different mechanisms: (1) neutralisation
of the carbonaceous, charged species by acids; (2) screening of the repulsive charge
effects by added salts and bases. Also, they used the Derjaguin, Landau, Verwey and
Overbeek (DLVO) theory for the van der Waals and electric double layer interactions
between bubbles, in a static, thermodynamic model. However, the main issue with this
interpretation must be based on the fact that the original and later work on bubble coa-
lescence inhibition was carried out with the specific exclusion of dissolved CO2 [5].

2.2.2. Simple experimental study to test the effect of dissolved CO2


Earlier bubble column coalescence studies used various gases not containing CO2
[5,31] but recent studies [2,3,13,15,16] were conducted using air. In both types of
experimental study bubble coalescence inhibition behaviour was similar, which indi-
cates that the presence or absence of dissolved CO2 was of little or no importance.
Although this is a review article, we are able to report here some simple experiments
on determining the influence of dissolved CO2 on bubble coalescence in 0.1 and 0.5 m
NaCl solutions in the BCE experiment.
Pure nitrogen was used as a carrier gas and room temperature was maintained within
the open-to-atmosphere bubble column and a pH probe was inserted to the centre of the
bubble column solution for the online pH measurement during the experimental runs.
The initial pH of the (CO2 equilibrated) electrolyte solution was measured at around 5.6.
After N2 bubbling for 2–3 min, the pH of the solution increased to slightly above 7, as
shown in Figure 3, which is the expected value. However, the pH continued to further
increase up to about 8.2. The additional increase observed must be caused by the vigor-
ous stirring of the high density of flowing bubbles within the BCE, which apparently can
set up streaming potentials, artificially increasing the apparent solution pH. Because of
this, pH values of solution samples taken from the BCE were also measured under N2
but in the absence of bubbles (as also shown in Figure 3) and these approach 7, as
expected. In separate experiments N2 sparging (without BCE), where bubbles were held
away from the glass pH electrode, did not show this artificial increase in pH (see
Figure 3). More importantly, as we can see from Figure 3, the bubble sizes and bubble
International Reviews in Physical Chemistry 149

Figure 3. (Colour online) Measured pH variation with time at room temperature using pure
N2 as a carrier gas measured within a BCE (red line) compared with pH measurements where
bubbles were prevented from contacting the pH glass electrode (black line). Also given are
photographs of the BCE at the start and after 100 min of bubbling with N2 gas.

densities in the photographs clearly demonstrate no difference between the beginning of


N2 gas sparging and after 100 min, where the pH increased to around 7, which further
supports the view that dissolved CO2 has little or no effect on coalescence.

2.2.3. DLVO interaction forces between bubbles


The expected DLVO interaction force between two bubbles of radius 1 mm immersed
in 1 M NaCl solution is shown in Figure 4. The Debye length of 0.3 nm reduces the
range of the electrical-double layer repulsion and hence the interaction forces are domi-
nated by attractive van der Waals forces. The surface potential of −40 mV was used
for this salt concentration, which actually corresponds to a much higher surface charge
density than expected. The van der Waals interaction was calculated using a non
retarded Hamaker constant of 37 × 10−21 J [33] which is the expected value for two
bubbles across water. This calculation demonstrates that at the high salt concentrations
where bubble coalescence inhibition is observed, the electrical double layer repulsion
between two bubbles is negligible and attractive van der Waals forces dominate. The
range of any, even weak repulsive forces cannot extend beyond about 1–2 nm, whereas
it is known that bubbles coalesce when the intervening film approaches about 50 nm
[40]. Hence, it is difficult to see how a DLVO based model could explain bubble coa-
lescence inhibition observed at very high salt concentrations.
Bubble coalescence and its inhibition are important phenomena with many applica-
tions, but there is still no realistic explanation for its occurrence [40].

2.2.4. Other factors affecting bubble coalescence or break up


2.2.4.1. Effect of 100% saturated bubble on coalescence. All of the earlier studies on
the effect of salt on bubble coalescence within bubble columns have used dry inlet
150 M. Shahid et al.

Figure 4. DLVO calculation of the force between two bubbles of radius 1 mm with a −40 mV
surface potential, Debye length of 0.3 nm, which corresponds to a concentration of 1 M for a 1:1
electrolyte, and a non-retarded Hamaker constant of 37 × 10−21 J, using the non-linear Poisson–
Boltzmann equation.

gases, which must cause the evaporation of water vapour into the bubbles formed at
the sinter. Consequently, an enhanced local salt concentration is expected near the sur-
face of the bubbles which could generate a substantial local, high osmotic pressure,
which would induce water to flow into the intervening film to cause a momentary
repulsive pressure acting to separate the bubbles. However, evaporation into the bub-
bles can be completely prevented by the use of inlet gas (air) pre-equilibrated with
water vapour to produce 100% relative humidity at the temperature of the bubble col-
umn. Recent experiments [3] (see Figure 5) did just this and compared vapour pre-equi-
librated gas with dried gas. Surprisingly, no visual difference was found in bubble
diameter and bubble density within the columns, at room temperature, in 0.5 m NaCl
solution. Hence, these simple experiments clearly demonstrate that this evaporation/os-
motic pressure model cannot explain the observed inhibition of bubble coalescence [3].

2.2.4.2. Effect of high temperature inlet gas on bubble coalescence. According to the
film drainage model, two colliding bubbles may be prohibited from combining by a
thin film of liquid trapped between them. Hence, the combined effect of an increased
column solution temperature and hot gas bubbles must, at least close to the sinter in
the BCE, produce significant local heating in the adjacent solution which should signif-
icantly reduce the film viscosity and hence might be expected to enhance bubble coa-
lescence. Again, however, no effect was observed in recent experimental studies using
hot gas bubbles in the range 150–275 °C. That is, the effect of added salt was found to
be the dominant factor in determining the degree of coalescence, rather than the effects
of increased temperature in the intervening film between approaching bubbles. Thus,
the effects of reduced water film viscosity can only be either absent or of negligible
importance (see Figure 6) [3,15].

2.3. Vaporisation model


A simple vaporisation model proposed in Garai’s work [46], improved for water and
salt solutions, described in Figure 7, offers a possible explanation for the observed
International Reviews in Physical Chemistry 151

Figure 5. Photograph of the equipment used to produce a flow of 100% relative humidity inlet
gas into the BCE.

Figure 6. (Colour online) Photograph of a bubble column containing 0.5 m NaCl solution with
inlet gas flow (dry air) at 275 °C (right photo) and pure water (left photo), for the same tempera-
ture and flow rate (adapted from Shahid and Pashley [15]).

Figure 7. Schematic cross section of the proposed vaporisation model at the liquid/vapour inter-
face (adapted from Fan and Pashley [13]).
152 M. Shahid et al.

variation in ΔHvap values with salt concentration. This model can be described by the
following equation [13]:
DHvap ¼ 8pr2 cNA þ RT þ DEHB PHB þ DEvdw ð1  PHB Þ (3)

where r is the radius of a water molecule; γ is the surface tension; NA is Avogadro’s


number; R is the universal gas constant; T is absolute temperature; PHB is the degree of
hydrogen bonding; ΔEHB is the energy required for completely breaking and separating
water molecules from within a tetrahedral arrangement i.e. with four hydrogen bonds
and ΔEvdw for the energy required to disaffiliate from attractive van der Waals forces.
The enthalpy of vaporisation has to be supplied for the internal energy to remove
the layer of surface, which is half the surface of the sphere (ΔUv = 2π(2r)2γNA, assum-
ing the water molecule is considered as a hard sphere and ignoring the length of the
hydrogen bond), and for the external work (expansion, ΔWv = PV/n = RT). However,
for water evaporation, it also involves intermolecular forces, such as hydrogen bonds
and van der Waals forces. For pure water, the surface energy term (ΔUv) is 0.28383γ
(with a radius of a water molecule of 1.37 Å [47]) which is very similar to the slope of
the fitted curve from the relation of literature surface tension values [48] with tempera-
ture from 0 to 40 °C, with a difference of only 0.65% [13], as shown in Figure 8.
Unfortunately, the behaviour of salt solutions is more complicated (i.e. non-linear)
because solvated ions have strong perturbations on the network of hydrogen bonds in
pure water [49]. However, the model can still give a relatively reasonable explanation
for the ΔHvap variations with concentration [13]. With the increase of concentration of
salt solution, the surface tension and van der Waals forces will increase and hydrogen
bonding decrease because of the perturbation of solvated ions. If hydrogen bonding is
the dominant factor, the value will decrease; whereas, the surface tension and van der
Waals energy would cause a net increase.

Figure 8. The change of ΔHvap with surface tension for different salt solutions at T = 298.15 K
and pure water at different temperatures (reproduced from Fan and Pashley [13]).
International Reviews in Physical Chemistry 153

2.4. Bubble evaporation layer model


The bubble evaporation layer model presented here is simply based on the calculated
thickness of a water layer around the surface of an initially dry bubble, which must be
evaporated into the bubble to produce saturated vapour pressure, at that temperature.
This surface bubble layer thickness (of water which must be vaporised) is a function of
temperature and bubble diameter and is useful in considering heat and mass transfer
processes for desalination, sterilisation and other applications. The bubble layer thick-
ness (tbl ) varies with temperature, and this model can be described by the following
equation for situations where the evaporated film thickness is much less than the radius
of the bubble:
1 qv ðTe Þ
tbl ¼ r w (4)
3 qw
where r is the radius of a bubble; qvw ðTe Þ is the water vapour density at the steady state
column temperature and qw is the liquid water density. Equation (4) shows that the
bubble layer thickness is a function of steady state column temperature (see Figure 9).

2.5. Bubble surface hot layer model


In addition to causing water evaporation around the surface of a hot, dry bubble enter-
ing a BCE, heat will also be transiently passed to a thin water layer surrounding the
bubble, as the bubble approaches steady state and cools to the temperature of the col-
umn solution. A transient hot water film will be produced around the bubbles flowing
into a BCE and this layer can be used to effect changes within the solution. It is there-
fore useful to consider the likely thickness of this heated layer as a function of inlet air

Figure 9. (Colour online) The relationship between the temperature of the inlet gas and the
calculated thickness of the water evaporation layer for a 1 mm radius bubble in pure water, 0.5 m
NaCl and 5.0 m CaCl2, required to evaporate to produce the saturated vapour pressure within the
initially dry bubbles (reproduced and adapted from Shahid and Pashley [15]).
154 M. Shahid et al.

temperature. These thermal effects appear to play a key role for heat transfer processes
like sterilisation [16–18] and solute thermal decomposition [2]. The maximum extent of
the layer can be roughly estimated for a given inlet air temperature, assuming that there
is a linear decrease in the temperature of the surrounding water film from the bubble
temperature to the column solution temperature, as illustrated in Figure 10. The thermal
energy supplied to heat the water film must be supplied by cooling of the freshly
released bubble. (Note that the millimetre bubble surface can be considered to be flat
relative to the thin, nanometre thickness of the heated water layer).
In this simple model we consider the intermediate state when the inlet bubble
(initially at Ti) has cooled from Ti to Tb and the heat transferred to the thin water film
surrounding the bubble is sufficient to produce this temperature profile in the film. Of
course, this amount of heat is the same as that required to heat a film of thickness δ
from the column solution temperature Te to the average film temperature (Tb + Te)/2. It
should be noted that the maximum bubble layer temperature will not be over 100 °C;
hence the average film temperature is estimated by assuming Tb = 100 °C. The bubble
layer thickness varies with bubble diameter, the temperature of the inlet gas and the
steady state column solution temperature. The maximum heated layer thickness can be
calculated using the thermal energy balance equation [2,16]:
Cp DT 0 V ¼ Cs Dt4pr2 qw d (5)
where Cp and Cs are the air and solution (or water) specific heat capacities respectively,
and qw is the liquid water mass density. Δt = (Tb + Te)/2 – Te and ΔT′ = Ti − Tb are the
transient temperature increase in the solution layer and the temperature reduction within
the cooling bubbles, respectively.
The volume of a layer of thickness δ around a bubble is given by 4pr2 d, when δ is
much smaller than r. Hence the cooling of the bubble by ΔT′ must determine the
thickness d. In addition, if we also assume that there is a 50% heat loss due to water
vaporisation then the thickness of the active hot region can be roughly estimated, as
shown in Figure 10. The shaded band in Figure 10, gives a rough guide to the expected
variation in hot water layer thickness, with the change of specific heat capacity due to
the change of salt concentrations.
This transient layer of heated water around the bubbles must be responsible for the
sterilisation effects [16–18] and solute thermal decomposition effects [2] recently
reported.

2.6. Bubble water vapour equilibration


It has been observed that fairly large air bubbles (1–3 mm) in water, which are used in
the BCE method, become non-spherical and oscillate both in shape and trajectory, thus
enhancing the rate at which water vapour equilibrates within the bubbles [50]. It is
remarkable that water vapour saturation within these bubbles is attained in a few tenths
of a second because of these oscillations and the circulatory fluid flow induced inside
the bubbles due to shear forces generated at the surface of bubbles which produces
rapid water vapour transfer within the bubbles [51], much faster than that expected for
quiescent diffusion, which would require several seconds to reach equilibrium accord-
ing to Fick’s law. This rapid vapour transfer must correspond to a similarly fast rate of
transfer of heat to the surrounding column solution and these factors form the basis for
several recent applications (see Figure 1) of the BCE process.
International Reviews in Physical Chemistry 155

Figure 10. (Colour online) Relationship between the temperature of the inlet gas and the estimated
thickness of the transient hot bubble surface layer around a 1 mm radius bubble in pure water, 0.5 m
NaCl and 5.0 m CaCl2. Specific heat capacity data of NaCl and CaCl2 solutions are from Clarke and
Glew [95] and Garrett [96], respectively. Density data of NaCl and CaCl2 solutions are from CRC
Handbook [48] (produced based on model and Equation (5) from Shahid et al. [2]).

2.7. Bubble rise velocity


Bubble rise behaviour in water, even for single isolated air bubbles, is surprisingly
complex [50,51] and depends on bubble diameter, sphericity and water purity [52]. The
presence of many other bubbles within a BCE makes this situation even more complex
and this has not been well studied. The rise of a bubble in a liquid is a function of
many parameters viz. bubble characteristics (diameter and shape), properties of gas–liq-
uid systems, and operating conditions, temperature, cleanliness etc.
It was demonstrated by Leifer et al. [50] that the motion of intermediate (single,
isolated) bubbles ranging from 1 to 3 mm diameter is produced by the combination of
two oscillation types, trajectory oscillations (zig-zag or helical), and shape or deforma-
tion oscillations (ellipsoidal). These gas bubbles actually rise at a limited rate of
between about 15 and 35 cm s−1 in quiescent water because they undergo trajectory
and shape oscillations which reduce their rise rate [53].
Quinn et al. [54] reported that the shape and velocity of ellipsoidal bubbles appears
to be linked to increasing solute concentration, creating more spherical bubbles with
reduced rise velocity and a unique bubble shape – rise velocity relationship, indepen-
dent of solute type.
It was also explained by Gonzalez-Tello et al. [55] that surfactants predominantly
modify the surface of a bubble through the adsorption of a monolayer producing a
more rigid interface and so enhancing fluid drag. The rise velocity in these solutions is
less than for clean bubbles of the same diameter.
Luo et al. [56] have studied the rise velocity of single bubbles in liquid-solid sus-
pensions at pressures up to 17 MPa and temperatures up to 88 °C, over the bubble
156 M. Shahid et al.

diameter range from 1 to 20 mm. It was found that the bubble rise velocity decreased
with increasing pressure and with decreasing temperature. The decrease of bubble rise
velocity was mainly due to the variations of gas density and liquid viscosity with pres-
sure and temperature.
Many researchers have described several factors affecting bubble rise velocity [52]
but the detailed understanding of bubble rise velocity and its associated parameters with
regards to the BCE method is yet to be thoroughly explained.
Stokes’s law was derived for spherical objects moving under high Reynolds number
and with zero slip boundary condition and gives rise rates substantially higher than
those observed for air bubbles in water. Unfortunately, the addition of slip boundary
conditions would give even higher rise rates, such as those obtained using the
Hadamard–Rybczynski (H–R) [57] equation (see Figure 11(a)). The Levich [58] for-
mula gives results more closely resembling experimental rates and this is also given in
Figure 11. Equation (6) represents the general formula for the theoretical bubble rise
calculation but unfortunately, none of these equations give an accurate explanation of
air bubbles in the 1–3 mm diameter range of interest [50,59]. The general formula is
given by the equation:
qw ga2
U1 ¼ k (6)
g

where U∞ is the bubble rise velocity in an infinite liquid, η is the coefficient of


viscosity of the liquid, qw is the density of the liquid, g is the gravitational constant, a
is the gas bubble radius and k is a constant (2/9 for Stokes equation, 1/3 for the H–R
equation and 1/9 for the Levich equation) (see Figure 11(a)). The three typical equa-
tions, discussed earlier, have closer agreement with the experimental results summarised
by Klaseboer et al. [60] in Figure 11(b) when bubble diameters are less than about
1.0 mm.
The complex behaviour of bubble rise rates for isolated bubbles in the
diameter range >1 mm will be further complicated by multiple collisions within a
densely packed bubble column, where coalescence is inhibited by the presence of
added salts.

Figure 11. (Colour online) The relationship between rise velocity of isolated bubbles and bubble
diameter using: (a) calculated values using Stokes’s law, H–R and Levich’s formula and (b)
experimental data and several typical models (published by Klaseboer et al. [60]).
International Reviews in Physical Chemistry 157

2.8. Thermal energy balance in the BCE


Consideration of the steady state thermal energy balance within a BCE, containing salt
solutions as illustrated in Figure 12, can be used to explain the process whereby the
heat supplied from the entering warm bubbles (per unit volume of gas leaving the col-
umn) is balanced by the heat required for vaporisation, to reach the equilibrium water
vapour pressure within these bubbles. This principle is based on the steady state volu-
metric balance within a bubble column, which has been used for the determination of
the enthalpy of vaporisation (ΔHvap) of concentrated salt solutions [3,4,13], and is
described by Equation (7):
 
DT Cp ðTe Þ þ DP ¼ qv ðTe Þ DHvap ðTe Þ (7)

where Cp(Te) is the specific heat of the gas flowing into the bubble column at constant
pressure; Te is the steady state temperature near the top of the column; ρv is the water
vapour density at Te, which can be calculated from the water vapour pressure of salt
solutions at the steady state temperature, using the ideal gas equation; ΔT is the temper-
ature difference between the gas entering and leaving the column; ΔP is the hydrostatic
differential pressure, between the gas inlet into the sinter and atmospheric pressure at
the top of the column, which represents the work done by the gas flowing into the base
of the column until it is released from the solution.
This equation was first published in 2009 by Francis and Pashley [4] for low
vapour pressure aqueous solutions, that is, at low column temperatures of about 283 K.
The equation’s accuracy and precision was further tested in later studies [3,13] at room
temperature. In a more recent variation of this energy balance equation, use of the heat
capacity per unit volume, Cp, in volume-based units was replaced by the corresponding
heat capacity per unit weight of gas Cpg (i.e. 1.005 J g−1 K−1 over 270–330 K [61] for
dry air), which is fairly constant with temperature. This gives a new version of the
thermal balance equation:

Figure 12. (Colour online) Schematic diagram of a bubble column.


158 M. Shahid et al.
h i
DT Cpg ðTe Þ qg þ DP ¼ qv ðTe Þ DHvap ðTe Þ (8)

where ρg is the density of air (or gas) which can be obtained using the molar mass of
air (28.96 g mol−1) where the absorbed water vapour is subtracted from the total num-
ber of moles of gas within the bubble per unit volume, using the ideal gas equation, at
Te and 1 atm pressure. It should be noted that the mass of air or gas within a bubble
remains constant as the bubble passes through the column. This equation appears to
encompass a more logical and reasonable thermo-physical energy balance to describe
the BCE process, especially at higher column solution temperatures, where the water
vapour pressure will be more significant.
This model assumes that while bubbles are capturing water vapour and rising in the
solution from their initial dry state to 100% water vapour saturation at Te, they will
expand further due to the water vapour captured into the bubbles but only the initial
(and constant) mass of gas can supply heat to the column, to produce this level of
evaporation.
In both balance equations ΔHvap includes the work done by vapour expansion.
However, for high gas temperatures and short bubble residence times, removing the
water vapour expansion work produces a better fit when using Equation (8) to give
Equation (9), which is adapted for no work of water vapour expansion and corresponds
to the following equation:
h i  
DT Cpg ðTe Þ qg þ DP ¼ qv ðTe Þ DHvap ðTe Þ  Patm DV ðTe Þ (9)

Hence, it appears that high temperature bubbles in short height column solutions
(~5 cm) and short residence times appear to show no expansion due to vapour uptake
(see Table 2). On the other hand, low temperature experiments appear to correspond to
a steady state condition in terms of not only temperature but also the bubbles’ vapour
expansion. It is interesting that at low temperatures, Equation (9) produces low errors
for calculated ΔHvap values, similar to those produced from Equations (7) and (8),
because the PatmΔV term, water vapor expansion work, has no significant effect on the
calculation. It should be noted that since bubbles reach vapour and temperature equilib-
rium within a few tenths of a second, the column height becomes important. In this
work we define a medium height column to be one where this equilibrium is just
attained.
It should be noted that use of Cvg , the heat capacity of the dry air under constant
volume conditions (0.718 J g−1 K−1 from 270 to 330 K [61]) in place of the corre-
sponding Cpg values gives inaccurate ΔHvap values at both high and low column solu-
tion temperatures. This indicates that atmospheric work done on the contracting,
cooling bubbles must be transferred to the column and so must be included in the
energy balance equations.
Based on the results obtained and the analysis of energy balances in the BCE, the
original balance Equation (7) was found to be accurate for the determination of ΔHvap
values of salt solutions over a wide range of temperatures. In comparison, Equation (8)
can only be used at low temperatures and Equation (9) is applicable to the special case
of high inlet gas temperatures and short bubble residence times. Although the original
Equation (7) was derived for low temperatures, around room temperature, we have
since shown that it also works well for very hot inlet gases [15]. This is because any
given bubble released into the column at high temperature (Tin) will contract until it
reaches the column solution equilibrium temperature Te. The heat supplied by the
Table 2. ΔHvap error percent using energy balance equations in BCE at high inlet gas temperature with and without surfactant C12EO8.

ΔHvap values
Inlet gas Column top kJ mol−1
temperature temperature
From From From Error percent Error percent Error percent
Solution type K K Literature Equation (7) Equation (8) Equation (9) Equation (7) (%) Equation (8) (%) Equation (9) (%)
0.5 mol kg−1 NaCl + 423.15 314.35 43.41 43.64 40.45 43.06 0.52 −6.83 0.80
0.002 g C12EO8 448.15 317.75 43.28 43.65 39.82 42.47 0.85 −7.98 1.88
473.15 320.75 43.16 43.73 39.26 41.93 1.31 −9.04 2.86
498.15 323.65 43.06 44.30 39.17 41.86 2.87 −9.04 2.80
523.15 325.25 42.98 45.46 39.66 42.37 5.77 −7.73 1.44
548.15 328.65 42.85 43.01 36.57 39.30 0.36 −14.66 8.28
0.5 mol kg−1 NaCl 423.15 314.85 43.39 42.32 39.15 41.77 −2.46 −9.78 3.75
448.15 317.75 43.28 43.65 39.82 42.47 0.85 −7.98 1.88
473.15 320.75 43.16 43.73 39.26 41.93 1.31 −9.04 2.86
498.15 323.55 43.05 43.59 38.47 41.167 1.26 −10.63 4.38
523.15 326.25 42.95 43.17 37.40 40.11 0.52 −12.92 6.60
548.15 327.75 42.89 44.99 38.53 41.25 4.91 −10.16 3.80
International Reviews in Physical Chemistry

Notes: All literature ΔHvap values were calculated from the literature vapour pressure values using Clausius–Clapeyron equation. NaCl solution vapour pres-
sure data were from Clarke et al. [95].
159
160 M. Shahid et al.

bubble (of volume Vb at temperature T) to the column, as it contracts, is given by the


sum of Cp(T) Vb(T) over the temperature range Tin to Te (i.e. ΔT). However, because
the mass of gas in any given bubble remains constant, the value Cp(T) Vb(T) is also
constant for any bubble and hence the heat transferred to the column is given by
Cp(Te) ΔT, per unit volume, as given in Equation (7).

2.9. Comparison of the BCE desalination process with multistage flash distillation
and reverse osmosis
Multistage flash distillation (MSF) is a very common desalination process but it is an
irregular process in that vapour transfer via boiling is less controlled than vapour col-
lection into a steady flow of dry air or gas bubbles. Boiling depends on nucleation,
which is hard to control and causes local temperature fluctuations resulting in energy
wastage. The enthalpy of vaporisation of water is about 2.3 MJ L−1 at 100 °C and
about 2.4 MJ L−1 at room temperature [51]. These values are not greatly affected by
the addition of salt. Although these values are high, most of this thermal energy is, in
practice, recycled on condensation of the water vapour and is used to heat the salt
water feed which reduces the overall energy cost. Commercial thermal/evaporative units
typically have energy costs in the range 20–200 MJ m−3, which reflects the high effi-
ciency of their heat recycling processes [62].
Reverse osmosis (RO) has become the most commonly used method in recent times
but it has many disadvantages which add to its inefficiency and cost. Large volumes of
concentrated saline retentate must be disposed of as only a fraction passes as clean
water through the membranes. The membranes themselves are expensive and easily
fouled and consequently, the saline solution must be pre-cleaned to protect the mem-
brane, adding to the complexity and cost of the process. The osmotic pressure of sea-
water is about 28 atm and high pressure liquid pumping is therefore required to
generate water through RO filtration and can be calculated with the following equation:
pos ¼ icRT (10)
where pos is osmotic pressure; c is concentration of seawater; R is gas constant; T is
absolute temperature; i is the van’t Hoff coefficient.
As the salt accumulates, the osmotic pressure required increases and commercially,
pressures up to about 70 atm are applied, which can increase the energy requirement to
about 10 MJ m−3 of water produced or more. The minimum work required to
desalinate seawater can be calculated from the work done by applying a pressure
infinitesimally higher than the natural osmotic pressure of the sea to obtain the reversi-
ble work done, at constant temperature, to move a semi-permeable membrane an
infinitesimal distance, so desalinating a very small volume of solution. This provides a
minimum work required of about 4.5 MJ m−3 of pure water, assuming concentrate at 2
× seawater concentration is produced. Commercial RO systems are less efficient,
typically in the range of 10–20 MJ m−3 [51]. The best-practice commercial energy cost
for the membrane desalination of seawater is currently at about 2.5 kW h m−3 or
9 MJ m−3 [62].
In contrast, BCE is a simpler process and easy to control, it does not produce scal-
ing as it operates at sub-boiling conditions, unlike MSF, thereby it is potentially a cost
and energy effective process. Also, BCE does not require feed water pre-treatment,
rather it is self-cleaning via the flotation process. In addition, BCE is suitable for
International Reviews in Physical Chemistry 161

Table 3. Theoretical water loss and energy consumption using BCE at 275 °C inlet gas temperature.

NaCl solution Vaporise 1 m3 No 99%


concentration in Theoretical water recovery recovery Energy
molality weight loss during required time electricity electricity utilisation
mol kg−1 25 min in g h kW h−1 kW h−1 MJ m−3
0.5 66.138 6299.95 755.99 7.55 27.21
1 64.5 6459.94 775.19 7.75 27.9
2 63.8 6530.82 783.69 7.83 28.21
3 65.47 6364.23 763.7 7.63 27.49
4 67.227 6197.9 743.74 7.43 26.77
5 62.861 6628.38 795.4 7.95 28.63
6 62.403 6677.02 801.24 8.01 28.84

continuous operation and produces high fresh water recovery rates to a much higher
concentration level of up to 6 m NaCl when compared to RO (see Table 3).
The results given in Table 3 were obtained using inlet air temperatures at around
275 °C and for solutions ranging from 0.5 to 6 m NaCl. The estimated results indicate
that the amount of water carried over varied with different NaCl concentrations relative
to the expected vapour carry over, obtained from the variable theoretical column tem-
perature (i.e. since the theoretical column temperature varies with increasing NaCl con-
centrations) and the corresponding vapour pressure. At the highest NaCl concentration,
6 m, the calculated energy utilisation at 275 °C was found to be about 29 MJ m−3
which indicates that the BCE process can be an energy effective method, especially at
higher solution concentrations and hence could be used for industrial applications.

3. Applications of the BCE


The gas–liquid direct-contact BCE is characterised by a continuously replenished, high
gas–liquid interfacial area, which subsequently offers higher mass and heat transfer
coefficients due to non-isothermal [7] (i.e. localised evaporation) nature of the BCE
process and could be applied in many large scale industrial applications.
The BCE currently has the following applications in aqueous systems:

(1) Determination of enthalpies of vaporisation (ΔHvap).


(2) Evaporative air conditioning systems for buildings.
(3) Seawater desalination.
(4) Water sterilisation.
(5) Thermolysis of different solutes in aqueous solutions.
(6) Inhibition of salt precipitation in concentrated salt solutions.

3.1. Precise measurements for the determination of enthalpies of vaporisation


(ΔHvap)
3.1.1. Background
The enthalpy of vaporisation (ΔHvap) is the amount of heat required to vaporise any
liquid in pure or solution form and it is therefore directly relevant to thermal desalina-
tion and evaporative cooling effects, as well as many other applications. ΔHvap of con-
162 M. Shahid et al.

centrated salt solutions is one of most important physical properties used in many
industries but literature values are quite variable. Classically, ΔHvap values are most
commonly determined through methods of calorimetric measurements. However, for
example, the ΔHvap results of pure water obtained by Lunnon [63] are quite variable
compared with CRC Handbook of Chemistry and Physics [48]. The BCE method
described here has given results in good agreement with literature ΔHvap values in
recent work, [13] quite contrary to the values reported by Gurovich et al. [64] which
are difficult to accept.
ΔHvap values also can be theoretically estimated using the Clausius–Clapeyron
equation (11), which is a fairly accurate way to obtain literature ΔHvap values, if there
are suitable literature vapour pressure-temperature values of salt solutions available.

  
DHvap 1 1
P2 ¼ P1 exp  (11)
R T1 T2
Use of this equation requires saturated vapour pressures, P1 and P2, at fairly similar
temperatures T1, T2, typically, say, no more than T = 5 K apart, since derivation of this
equation is based on the assumption that ΔHvap is fairly constant over this temperature
range; R is the gas constant.

3.1.2. Determination of ΔHvap values for concentrated aqueous solutions


Bubble coalescence inhibition and rapid vapour transfer both offer a novel application
of the BCE process for the precise measurement of ΔHvap values for concentrated salt
solutions [3,4,13] where the vapour pressure values of salt solutions are known. ΔHvap
values can be obtained from the volumetric energy balance within the column, operat-
ing under steady state conditions, as defined by Equations (7) and (8) (in Section 2.8).
In recent work [3,13], ΔHvap values of various salt solutions at different concentrations
have been obtained with accuracies, on average, within around 0.5–1.0% compared to
the literature values for several salt solutions and demonstrate a measurement precision
of between 0.1 and 0.9%. For relatively low column temperature studies, around room

Figure 13. (Colour online) Normal distribution diagrams of experimental ΔHvap values at
T = 298.15 K for the common salt solutions using the data acquisition system: (a) 1.0 m NaCl
solution and (b) 5.0 m CaCl2 solution (adapted from Fan and Pashley [13]).
International Reviews in Physical Chemistry 163

Table 4. Comparison of ΔHvap values obtained using the BCE method in the range between 19
and 26 °C for several salts.

ΔHvap values
kJ mol−1
Error percent Error percent
Steady state Experiment Experiment from from
temperature using using Equation (7) Equation (8)
Solutions °C Literature Equation (7) Equation (8) (%) (%)
NaCl (1.0 m) 21.648 ± 0.010 44.17 44.89 44.23 1.6 0.2
21.598 ± 0.040 44.17 44.48 43.38 −0.2 −1.8
21.120 ± 0.013 44.19 44.78 44.04 1.3 −0.3
21.121 ± 0.007 44.19 44.53 43.94 0.8 −0.6
24.996 ± 0.178 44.04 43.43 43.06 −1.4 −2.2
26.375 ± 0.093 43.98 43.62 43.07 −0.8 −2.1
CaCl2 (1.0 m) 20.428 ± 0.022 44.16 45.49 43.79 0.7 −0.8
19.769 ± 0.078 44.19 45.52 44.81 3 1.4
25.378 ± 0.233 43.97 44.14 43.67 0.4 −0.7
25.701 ± 0.144 43.96 44.31 43.80 0.8 −0.4
25.796 ± 0.160 43.96 44.25 43.74 0.7 −0.5
CaCl2 (5.0 m) 20.206 ± 0.081 45.65 45.97 45.57 0.7 −0.2
20.487 ± 0.008 45.65 45.37 44.95 −0.6 −1.5
20.566 ± 0.016 45.64 46.26 45.80 1.4 0.3
25.896 ± 0.187 45.50 45.74 45.35 0.5 −0.3
26.135 ± 0.089 45.50 45.88 45.48 0.9 −0.03
KCl (4.0 m) 20.835 ± 0.037 43.16 44.14 43.59 2.3 1.0
20.336 ± 0.023 43.17 44.32 43.72 2.7 1.3
21.026 ± 0.049 43.16 43.71 43.18 1.3 0.04
Notes: All literature ΔHvap values were calculated from the literature vapour pressure values
using Equations (7) and (8). NaCl solution vapour pressure data were from Clarke et al. [95];
CaCl2 and KCl solution vapour pressure data were from Patil et al. [97]. It should be noted that
even for literature ΔHvap values, they have variations for common salt solutions [3] (data repro-
duced from Fan et al. [3]).

temperature, the two balance equations, discussed earlier, give similar results, i.e.
within the 0.5–1.0% range, on average, as shown in Table 4.
The accuracy and precision of these ΔHvap measurements was recently further
improved using vacuum insulation (which reduced thermal transfers with the room
environment) of the bubble column and an automatic data acquisition system for tem-
perature readings [13]. Under the steady state of the BCE system, hundreds of tempera-
ture data can be acquired by computer within 0.5–1 h and hence a large number of
ΔHvap values can be produced using these energy balance equations. These ΔHvap val-
ues [13] were found to pass the Kolmogorov–Smirnov test [65] with high probabilities
after standardisation of data and supported the null hypothesis, that is the normal distri-
bution. Typical examples are shown in the frequency histograms given in Figure 13
[13]. Narrow confidence intervals (95%), say always around ±0.02 kJ mol−1, also
demonstrated the high precision of the BCE method for ΔHvap measurement, as well as
supporting the basic physics of the energy balance equation.
At relatively higher operating or steady state temperatures, around 40–50 °C, the
BCE method was still found to be suitable for determining ΔHvap values of the salt
solutions, for example, see the earlier data presented in Table 2 for 0.5 m NaCl aqueous
solutions.
164 M. Shahid et al.

3.2. Evaporative cooling


3.2.1. Background
The cooling effect produced when a liquid changes to a gas or vapour is known as
evaporative cooling. Two common household items – the clothes dryer and the evapo-
rative cooler – clearly demonstrate the thermal energy requirement which is substantial
for water evaporation. In the latter case, energy stored in the hot, dry, outside air is
consumed by the evaporation process as the air passes through or across the evaporat-
ing liquid. This evaporation reduces the temperature of the liquid solution and allows
us to use evaporative cooling. BCEs can be used for evaporative cooling processes with
concentrated aqueous salt solutions which inhibit bubble coalescence and so produce
high density bubble columns. Dry, warm air passed through the column will be contin-
uously cooled by this evaporation [4].

3.2.2. BCE evaporative cooling


The steady state operating temperature of an aqueous solution in a bubble column can
be calculated using the volumetric energy balance equation (7). The relation between
inlet gas temperature and column top temperature using known ΔHvap values and
vapour pressure values for salt solutions or pure water, can be calculated as shown in
Figure 14, for pure water, 0.5 m NaCl and 5.0 m CaCl2 solutions. The steady state col-
umn solution temperature within the BCE is a function of the temperature of the inlet
gas [3,4]. This cooling effect has led to the suggestion that the BCE process could be
used as a simple evaporative cooling system for buildings which was earlier proposed
by Francis and Pashley, published in 2009 [4].
When the BCE process is used with some types of salt solutions, which have coa-
lescence inhibition effects on the bubbles [31], it will produce a high volume fraction
of small bubbles (see Figure 6) which will enhance the water vapour transfer into the
bubbles. Figure 15(a) [4] shows that the solution with dissolved salt (NaCl) at 0.5 M
has a more efficient evaporative cooling effect and hence halved the time for the

Figure 14. (Colour online) Calculated steady state temperatures for a bubble column containing
pure water, 0.5 m NaCl and 5.0 m CaCl2 solutions as a function of inlet (dry) gas (air) tempera-
ture under atmospheric pressure (adapted from Fan et al. [3]).
International Reviews in Physical Chemistry 165

column to reach steady state conditions. Another important factor is the gas flow rate.
The higher the rate that gas is pumped in, the higher rate of cooling and the faster the
column will reach the steady state temperature, as shown in Figure 15(b). For an evap-
orative air conditioner system, for example, gas pumped in at 30 °C under a rate of
98 m3 h−1 would, with the column steady state temperature of 12 °C, produce
2000 BTU h−1 of cooling or 27 L s−1 of 12 °C moist air [4].
An indirect evaporative cooling method which delivers dry, cool air can be
designed as shown in the schematic diagram in Figure 16. In this system a metal coil,
as a heat exchanger, is placed within the bubble chamber and bubbles are pumped into
and then leave this cooler; the warm air therefore can be piped into the heat exchanger
to provide a continuous supply of dry, cool air [4].

Figure 15. (a) difference between pure water and 0.5 M NaCl solution in time taken to reach
steady state temperature, with both flow rates at 0.18 L s−1 and with inlet gas temperature around
22 °C and (b) difference in time taken to reach steady state temperature in BCE containing 0.5 M
NaCl solution with varying gas flow rates from 0.03 to 0.15 L s−1 and inlet gas temperature
around 25–27 °C (adapted from Francis and Pashley [4]).
166 M. Shahid et al.

Figure 16. (Colour online) Schematic diagram of BCE process for a cooling system (adapted
from Francis and Pashley [4]).

3.3. Seawater desalination


3.3.1. Background
Ancient processes of desalination were based on simple distillation and filtration [66].
Modern desalination processes can be divided into two main categories: (a) thermal
methods, which involve heating water to its boiling point to produce water vapour; and
(b) membrane processes, which employ a membrane to move either water or salt to
induce two zones of differing concentrations to produce fresh water [67]. Commer-
cially, the two most important technologies are based on the MSF and RO processes.
For example, in 1999 approximately 78% of the world’s seawater desalination capacity
was made up of MSF plants, while RO represented 10%. However, there has been a
gradual increase in RO seawater desalination primarily due to its lower cost and sim-
plicity [68].
In 2003 Fiorenza et al. [69] claimed that the typical average capacity and corre-
sponding cost for desalination technologies worldwide was:

• MSF: 25,000 m3 d−1 and 0.88 € m−3 (1.10 $ m−3)


• MED: 10,000 m3 d−1 and 0.64 € m−3 (0.80 $ m−3)
• VC: 3000 m3 d−1 and 0.56 € m−3 (0.70 $ m−3)
• RO: 6000 m3 d−1 and 0.56 € m−3 (0.70 $ m−3)

Thermal methods are more expensive because of the large quantities of fuel
required to vaporise salt water [70]. Membrane methods, mainly RO, which can desalt
brackish water somewhat more economically, have replaced thermal methods for the
desalination of brackish water. However, because of the high cost of membrane replace-
ment and less margin for improvement and need for extensive pre-cleaning of the feed-
water, membrane methods could be further improved for seawater desalination.
Technological advancement in thermal methods has also allowed some reduction in the
total desalination cost, by improving energy efficiency (MSF or hybrid systems or other
innovative technologies), by facilitating transfer processes or by energy recycling
International Reviews in Physical Chemistry 167

(process of cogeneration) [71] and bubble column desalination is potentially one of


these innovative and emerging thermal technologies.

3.3.2. Bubble column desalination


The dynamic BCE method has higher overall mass transfer and heat transfer coeffi-
cients, compared with quiescent systems [1]. In this novel bubble column desalination
system, water vapour can be captured and transported using a simple BCE system oper-
ated at temperatures well below the boiling point. The inhibition of bubble coalescence
in salt solutions enables the design of a bubble column with a high volume fraction of
small air bubbles, continuously colliding but not merging. This produces a uniform and
efficient exchange of water vapour into the bubbles, which together with the high bubble
rising velocity, due to its shape and trajectory based oscillations, allows water vapour to
be rapidly absorbed into bubbles, condensed and then collected as pure water [14].
The BCE method for seawater desalination, was examined and patented in 2013
[62]. The process is, of course, a reduced version of the natural phenomenon in which
air is used as a carrier gas for desalting seawater through the rain cycle. However, the
BCE process is based on the unexpected property of seawater in stopping air bubble
coalescence [5,31] because this facilitates a high packing volume of air bubbles, which
are persistently colliding but are prevented from coalescing by the presence of salts. In
addition, the bubbles produced in the 1–3 mm diameter range are ideally suited for
rapid water vapour uptake [50]. These factors form the basis of this enhanced process
for the desalination of seawater. In addition, the BCE process offers an efficient vapour
transfer mechanism in a continuous flow, evaporative bubble column operating below
the boiling point. This method provides a very high surface area of air–water interface
continuously produced and managed, naturally, by gas bubbling in salt water, such as
seawater, to improve the efficiency of evaporation and transportation of the water satu-
rated vapour producing drinking water from seawater [51].
In the first reported experiments [14], seawater was first heated to 70 °C then air
bubbles ranging from 1 to 3 mm diameter were produced via a glass sinter (porosity
No. 2) and was passed continuously through heated seawater. After bubbling for
60 min, the temperature of the solution in the column had fallen to about 52 °C. The
starting and finishing temperatures were used to estimate the theoretical yield expected
for complete collection and condensation of the water vapour, at the average tempera-
ture of the column. The electrical conductivity of the bubble column solution was
reduced from 49 mS cm−1 (seawater) to 6 μS cm−1 (well below the levels required for
acceptable drinking water) [14].
In a second method, both evaporation and condensation methods were controlled
within the same column, by maintaining a substantial temperature differential across
the column. Seawater was preheated to give a starting temperature of 65 °C and then
air flow at room temperature was continuously bubbled into the column. The top sec-
tion of the glass column was cooled using a glass condenser which had a cooling water
inlet at a temperature of 0.173 °C and an outlet temperature of 6.112 °C. Accordingly,
after several hours, this column of medium height and with a relatively small tempera-
ture differential of only just over 20 °C, produced a significant concentration gradient
of 0.6 M NaCl at the base to 0.15 M NaCl at the top. This dilution rate, of about 4
times, was very close to that expected from the estimated amount of water vapour
carried over for the flow rates and temperatures used. One of the advantages is that
renewable sources of energy (i.e. solar energy system) could easily be coupled with a
168 M. Shahid et al.

Figure 17. (Colour online) Schematic diagram of bubble column desalination process coupled
with a black pipe solar heater system (reproduced from Francis and Pashley [14]).

bubble column system [14]. It is also pertinent to mention that the BCE process could
possibly be coupled with hot industrial waste flue gases, as shown in the schematic
diagram in Figure 17.

3.3.3. Improved bubble column desalination


In some recent studies [3,15,72], high temperature inlet gases have been used to evalu-
ate the potential for improvement of the BCE process for seawater desalination. The
bubble column can actually be heated solely by the use of very hot inlet gases which
are often available industrially as waste. Enhanced bubble column desalination studies
were conducted using both steady state and supersaturated conditions.

3.3.3.1. Steady state bubble column desalination. In these experiments hot, dry, air bub-
bles at temperatures between 150 and 250 °C were passed through empty columns at flow
rates of about 23 L min−1 and then a known mass of solution (200 g of 0.5 mol kg−1
NaCl) was then quickly added. The temperature of the solution was then measured every
minute throughout each 30 min bubbling run. After 30 min, the column and remaining
solution was detached and weighed. Since the dry weight of the column was also known,
the total amount of water vapour removed in each experiment was easily measured. Typi-
cal results obtained showed 6.2% and 8.3% increased vaporisation by using 150 and
250 °C inlet gas temperatures, respectively [3,72], which indicate that increasing the inlet
air temperature can induce greater supersaturation levels in the bubbles.

3.3.3.2. Enhanced supersaturated bubble column desalination. In these experiments


hot, dry, air bubbles ranging from 150 to 275 °C at about 23 L min−1 were first passed
through an empty dry BCE column and then a known mass of the solution was quickly
added. In these experiments 200 g of 0.5 m NaCl was added with and without 0.002 g
of a non-ionic surfactant (C12EO8). The temperature of the solution was then measured
International Reviews in Physical Chemistry 169

Figure 18. (Colour online) Enhanced percent vaporisation as a function of gas (dry air) inlet
temperature for columns containing 0.5 m NaCl solution, with and without added C12EO8
surfactant. The percent was calculated by a comparison with the vaporisation expected from the
equilibrium vapour pressures at the operating temperatures of the bubble column measured during
the 30 min runs (reproduced from Shahid and Pashley [15]).

every minute throughout the 30 min bubbling runs. After 30 min, the column and
remaining solution was detached and weighed. Since the dry weight of the column was
also known, the total amount of water vapour removed in each experiment was easily
measured (see Figure 18) [15]. The results obtained show that increasing the tempera-
ture of the inlet air increases the water vapour carryover expected from the column
solution vapour pressure and the air volume passed. Further, the results also show that
with added surfactant the carryover increases even more.
It seems likely that the use of the non-ionic surfactant, octaethylene glycol monodo-
decyl ether (C12EO8), provides a monolayer coating at the surface of bubbles, as
described schematically in Figure 19.

Figure 19. (Colour online) Schematic diagram of adsorbed C12EO8 molecules (reproduced from
Shahid and Pashley [15]).
170 M. Shahid et al.

Figure 20. (Colour online) Measured solution temperatures within a bubble column containing
0.5 m NaCl solution with 0.002 g of added surfactant (octaethylene glycol monododecyl ether)
over 30 min, for 275 °C inlet gas (dry air). The measured weight of water evaporated was found
to be 83.9 g and the expected weight from the column temperatures was calculated to be 72.8 g.
By comparison, the theoretically expected weight (that is, with the column operating at thermal
equilibrium) was 72.2 g (reproduced from Shahid and Pashley [15]).

Thus, it appears that the packed mono-layer of surfactant molecules allows water
vapour transport into the bubbles but inhibits this vapour from re-condensing on the
interior, now hydrophobic, walls of the bubbles. Hence, the surfactant layer acts like a
‘surface molecular diode’, which facilitates water vapour transport in one direction.
This supersaturation of the air bubbles then produces increased water vapour carryover
[15] (see Figure 20).

3.3.4. Comparison and benefits


The main advantages of the BCE desalination system are its simplicity, resilience to
feed water purity and the fact that it is a continuous and controlled, non-boiling pro-
cess. These are clear advantages over the two most common seawater desalination pro-
cesses currently used, that is seawater reverse osmosis (SWRO) and thermal
desalination (such as MSF). There is little room left for improvement in SWRO but
thermal desalination methods can still be substantially improved. Methods such as the
BCE which represent low capital investment do not rely on rare materials or complex
manufacturing and ready use of waste heat and sustainable energy sources, such as
wind power, offer fewer constraints than the other common processes.
The main advantages in the BCE system for water desalination are listed below.

• The BCE collects water vapour throughout the entire body of the salt solution as
compared to MSF, which uses only the surface of the heating plates as the main
water vapour transfer site.
• The BCE process is uniform and controlled because it does not involve boiling.
• Very fast vapour collection of a few tenths of a second for 1–3 mm diameter
bubbles.
International Reviews in Physical Chemistry 171

• High rising velocity of saturated bubbles within the BCE system.


• Air flow produces continuously renewed bubbles and high surface area of evapo-
rating surfaces.
• No requirement for feedwater pre-treatment as in SWRO.
• System is self-cleaning via the flotation process.
• Sub-boiling process easier to control compared with MSF and doesn’t produce
scaling.
• Can concentrate to a much higher level, up to 6 m NaCl, than either SWRO or
MSF, i.e. produces higher recovery rate.
• Simple design will give low capital cost compared with MSF.
• Single stage and continuous process produces high quality water.
• Well suited for sustainable wind power for air flow generation and use of waste
industrial hot gases.
• Does not need vacuum pumps to reduce pressure and boiling point, compared
with MSF.
• Use of heated gas inflow offers an ideal process for control of column tempera-
ture and hence evaporation rate.

3.4. Water sterilisation


3.4.1. Background
Sterilisation is a physical or chemical process that destroys, eliminates or inactivates all
micro-organisms and bacterial spores. Conventional sterilisation techniques rely on irre-
versible metabolic inactivation or on the breakdown of vital structural components of
the micro-organisms. Micro-organisms can be destroyed (irreversibly inactivated) by
established physical microbiocide treatments, such as heating, UV or ionising radiation,
or by methods of new non-thermal treatments such as high hydrostatic pressure, pulsed
electric fields, oscillating magnetic fields or photodynamic effects; or a combination of
physical processes, such as heat–high pressure or heat-irradiation. Mechanical removal
of micro-organisms from food may be accomplished by membrane filtration of food
liquids [73]. Sterilisation is fundamentally categorised into thermal and non-thermal
methods and bubble column sterilisation offers a novel, energy-efficient thermal method
for sterilising water in sub-boiling conditions.

3.4.2. Bubble column sterilisation


The high heat transfer coefficients created within the BCE system can be used to ther-
mally destroy biological organisms well below the boiling point and it has recently
been established that sterilisation occurs due to transitory impact of biological species
with hot gas bubbles by collisions with the heated air–water interface, although the col-
umn temperature remains low and actually even favours the growth of bacterial colo-
nies present in typical contaminated water [17,18].

3.4.2.1. Bubble column faecal coliform inactivation using high inlet gas temperatures.
In the first BCE sterilisation study [18], hot gas bubbles up to 150 °C were used and in
a more recent study bubbles up to 250 °C [17] were passed into a water column via a
glass sinter with 40–100 μm pores. The effects of exposure on sterilising water were
examined with different time intervals and typical results are summarised in Table 5.
172 M. Shahid et al.

Table 5. A summary of column sterilisation experiments using the total column solution for
coliform counting (adapted from Shahid [17]).

Run time
min
Inlet gas temperature
°C 1 2 15 20 30 40 60 70
80 HFC HFC HFC HFC MFC MFC MFC MFC
90 HFC HFC HFC HFC MFC MFC MFC LFC
100 HFC HFC HFC MFC MFC LFC LFC ANFC
150 HFC HFC MFC LFC LFC LFC ANFC ANFC
200 MFC LFC LFC ANFC ANFC ANFC ANFC ANFC
250 LFC ANFC ANFC ANFC ANFC ANFC ANFC ANFC
Notes: HFC is high faecal coliform numbers; MFC is medium faecal coliform numbers; LFC is low
faecal coliform numbers; ANFC is almost no faecal coliform numbers, i.e. >99.99% sterilisation.

The degree of sterilisation was determined using natural lake water heavily contami-
nated with coliforms from waterfowl and land run-off. These coliform counts, obtained
through the membrane filtration method [74], were used to measure the degree of steril-
isation using hot gas bubbling under a range of conditions but in all cases where the
column solution temperature never exceeded the optimum growing temperature for the
coliforms [17,18].
The presence of a salt that inhibits bubble coalescence in the solution serves to pre-
serve finer bubbles of the heated gas, enhancing the number of bubble collisions with
the biological species by ensuring a higher gas–liquid contact surface area and higher
surface area per gas volume which leads to improved sterilisation rates (see Figures 21
and 22).
Figure 21 (and Figure 22) illustrates that the use NaCl enhances the rate of water
sterilisation as air bubbles more typically in the approximate size range of 1–3 mm
diameter increase the probabilities of bubble collisions with typically small microorgan-
isms ranging from 2 to 5 μm in size. It is worth mentioning that after 2 min very few
(blue) coliform colonies were still observed for columns with added NaCl in the solu-
tion, which produced the smaller bubbles.

Figure 21. (Colour online) Sterilisation rate as a function of an inlet gas temperature of 250 °C
for 2 min, for columns containing lake water samples, with (left photo) and without (right photo)
added 0.175 m NaCl. After 2 min of bubbling with smaller (mm size, with added salt) hot
bubbles, complete sterilisation was observed, whilst the same time is insufficient for the relatively
larger (cm size) hot bubbles (reproduced from Shahid [17]).
International Reviews in Physical Chemistry 173

Figure 22. (Colour online) Faecal coliform densities with time for (sinter type 2) bubble columns
containing 250 mL raw lake water with 0.01 M CaCl2, 0.15 M NaCl and without added salts,
using 150 °C dry air inlet at a flow rate of 25 L min−1 (reproduced from Xue and Pashley [16]).

3.4.2.2. Bubble column faecal coliform inactivation with added CaCl2. The coliform
inactivation results summarised in Figure 22 show that adding 0.01 M CaCl2 has a sig-
nificant effect on the inactivation rate and is more effective than without added salt,
increasing the inactivation rate from approximately 29.0% to 99.9% during 40 min
treatment. The results suggest that adding CaCl2 produces more effective and efficient
faecal coliform inactivation, due to its effect on reducing the repulsive surface forces
between the negatively charged bubbles and coliforms [16], since at this concentration
CaCl2 does not prevent bubble coalescence and so the bubbles diameter would be quite
similar to those observed with raw lake water.
It is recognised that energy is one of the key factors required for social, economic
and particularly industrial development [75] and the worldwide use of thermal sterilisa-
tion (i.e. boiling) of water suggests that substantial energy savings are possible by
adoption of the BCE process. The thermal energy cost to produce water sterilisation by
passing 23 L min−1 of air at 150 °C for 60 min and 250 °C for 2 min, through a suit-
able bubble column, can be compared with the standard process of sterilisation through
simple boiling. For 1 L of water it was calculated that the thermal energy supplied in
the first study was 150 kJ L−1 and for the second improved study was only 64 kJ L−1,
compared with about 450 kJ L−1, which was needed for the standard process of
sterilisation by boiling the contaminated water for 5–30 min, as typically recommended
[17,18].

3.4.3. Proposed mechanism of BCE sterilisation


In the BCE process, the initially hot, dry air bubbles must produce a thin layer of
heated water around the surface of the bubbles, as already discussed in Section 2.5.
The transient collisions between these heated water layers, as well as the hot air surface
of the bubbles, and the biological organisms dispersed in solution appear to be the
174 M. Shahid et al.

fundamental mechanism of coliform inactivation via thermal/biological changes in the


microorganism.
Rising hot bubbles within the aqueous BCE must initially and transiently produce a
thin (nms) hot layer of water around the surfaces which were found to thermally sterilise
biological organisms well below the boiling point [16–18]. This hot bubble layer when
colliding with a microrganisms’ membrane appears to cause thermal disruption and con-
sequently the destruction of the biological membrane resulting in thermal inactivation.
The sterilisation model presented here can be principally described by three inde-
pendent sub-steps: (1) particle-bubble collision; (2) particle-bubble attachment; and (3)
particle destruction.
Est ¼ Ec ! Ea ! Ed (12)
where, Est, Ec, Ea and Ed are the efficiency or probability of sterilisation, collision,
attachment and destruction of undesired biological organisms, respectively.
In this model, the destruction of Escherichia coli (E. coli) was shown in three simple
steps; the first step depicts the collision of 2–5 μm biological organisms with 1–3 mm
air bubbles, the second step shows particle-bubble attachment and lastly, the destruction
of cell membrane due to the interaction with transient hot rising air bubbles.
In Table 6, BCE water sterilisation was compared with the use of normal boiling.

3.5. Thermolysis of solutes in aqueous solution


3.5.1. Background
Thermal sterilisation studies with the BCE led to the suggestion that this process could
also facilitate thermal decomposition of some solutes in aqueous solutions, even at
lower solution temperatures and at a faster rate than is normally produced via the direct
heating of a bulk solution. Studies of the use of the BCE process for the thermal
decomposition of solutes have recently been reported [2,76]. One study examined the
thermal decomposition of ammonium bicarbonate (NH4HCO3) in aqueous solution.
This salt has been used for important applications, such as a draw solution in Forward
Osmosis [77] and, more recently, in the regeneration of ion exchange resins [78]. A
second solute in widespread use is potassium persulfate (K2S2O8) which was also stud-
ied in aqueous solutions. This salt that is often used as a radical initiator for the process
of emulsion polymerisation [79]. A schematic diagram of the BCE process for thermal
decomposition using a hot air bubble layer is shown in Figure 23.
Table 6. Comparison of water sterilisation by boiling and using bubble column method (adapted
from Shahid et al. [18]).

Sterilisation with boiling Bubble column sterilisation


Irregular process which is hard to control resulting in Controlled, regular process.
energy wastage
Produces fouling/scale formation Self-cleaning method
Requires a considerable amount of energy Significantly lower thermal energy usage
Does not remove suspended or dissolved compounds Flotation helps to remove suspended and
dissolved compounds
Potential for burn injuries Safe, low temperature method
Indoor pollution: increased risk of respiratory Reduced health risk by use of a sealed
infections from indoor stoves or fires inline gas electrical heater
After boiling, water needs cooling down (cannot be No need to cool
consumed immediately)
International Reviews in Physical Chemistry 175

Figure 23. (Colour online) Schematic diagram of BCE thermal decomposition using a hot air
bubble layer (adapted from Shahid et al. [2]).

Figure 24. (Colour online) Decomposition percent of NH4HCO3 solutions at different concentra-
tions with and without sucrose in the BCE (with an inlet air temperature of 150 °C and column solu-
tion temperature of 45 °C) and in a water bath at around 45 °C (adapted from Shahid et al. [2]).
176 M. Shahid et al.

3.5.2. Decomposition of NH4HCO3 solutions


Some typical decomposition results obtained using different solution conditions [2] are
given in Figure 25. These results clearly demonstrate that the BCE process is much
more efficient for NH4HCO3 decomposition, especially compared with the standard
method, which is, using a stirred water bath at the same solution temperature, of 45 °C.
Furthermore, different concentrations of NH4HCO3 decomposition were also studied
[2] as shown in Figure 24.

Figure 25. Photographs of the bubbles diameter in NH4HCO3 solution with corresponding con-
centrations at different experimental times: (a) 10 min at around 1.3 m; (b) 30 min at around
0.09 m; and (c) 50 min at around 0.01 m (adapted from Shahid et al. [2]).
International Reviews in Physical Chemistry 177

The initial high (2 m) concentration of NH4HCO3 used in these BCE experiments


was found to inhibit bubble coalescence, producing small bubbles. However, after
30 min of the BCE process, the significant reduction in NH4HCO3 concentration pro-
duced larger bubbles of the same diameter as those observed in pure water, which also
confirms the decomposition of NH4HCO3 salt into ammonia and carbon dioxide gases
[2], as shown in Figure 25.

3.5.3. Decomposition of K2S2O8 solution using a BCE process


In a similar series of experiments, the thermal decomposition of 0.05 m K2S2O8 solutions
was studied, with and without added sucrose (0.5 m), in a BCE process with inlet dry air
at about 150 °C. For comparison, decomposition was also studied using a continuously
stirred water bath. Both the water bath and BCE solution had an operating temperature
around 47 °C, for direct comparison. As K2S2O8 has a reasonably low solubility in water,
of around 0.16 m, the BCE process was designed using added sucrose in order to main-
tain finer bubbles for more efficient decomposition [2] (see Figure 26), since it has been
reported that sucrose also inhibits bubble coalescence [5,31].
It is worth noting that traditional decomposition of ammonium bicarbonate
(NH4HCO3) and potassium persulfate (K2S2O8) needs substantial thermal energy input
and studies show that the decomposition of ammonium bicarbonate solution occurs
between 30 and 85 °C [80,81]. Similarly, the decomposition of potassium persulfate
(K2S2O8) in solutions appears between 50 and 90 °C [82]. However, using the BCE
process ammonium bicarbonate and potassium persulfate were completely decomposed
at a solution temperature of only around 45 °C.

3.5.4. Proposed mechanism of BCE solute thermolysis


Pre-heated gas bubbles introduced and passed though the aqueous solution, must pro-
duce a transient hot surface layer around each rising bubble. The transient hot surface

Figure 26. (Colour online) Percent of K2S2O8 solution decomposition using a BCE process (with
inlet dry air at a temperature of 150 °C) and a water bath, both at around 47 °C, with and with-
out sucrose in the BCE process and water bath with added sucrose only (adapted from Shahid
et al. [2]).
178 M. Shahid et al.

layer will have a higher temperature than the average temperature of the aqueous solu-
tion and it is believed that it is the interaction of the solute with this transient hot sur-
face layer which results in the thermal decomposition of the solute, even when the
average temperature of the aqueous solution remains below the temperature at which it
would normally cause thermal/chemical decomposition of the solute. For situations
where thermal decomposition is either required very quickly or at reduced temperature,
the BCE method offers a new approach.

3.6. The controlled growth of precipitates in aqueous salt solutions


3.6.1. Background
Calcium sulphate (CaSO4) or calcium sulphate dihydrate (CaSO4·2H2O) are common
minerals precipitated from seawater and natural brines, and in practice, they are also
often a problem in the form of scale deposits produced in many industrial processes,
such as in water treatment and desalination [83]. These precipitates have been
thoroughly studied for precipitation kinetics [84–87], morphology and spectral
characterizations [88–90]. The combination of uniform bubbles, rapid water vapour
transfer and vigorous mixing with uniform solute concentration within a BCE process
can be used as the basis for a novel approach to produce precipitated particles over a
wide range of size in a controlled manner. Also, the BCE process has the possibility of
several flexible control variables, such as flow rate, gas type, gas humidity and inlet
gas temperature [4,13] to help produce optimum conditions. The BCE method was
found to have a significant inhibition effect on precipitation rate, which facilitates the
collection of particles over a wide range from nanometre to micrometre size.

3.6.2. Controlled CaSO4·2H2O precipitation in a BCE and comparison with simple


stirring
At first, it might appear that the BCE process, with continuous water evaporation via
the rising dry bubbles, could be used to slowly increase supersaturation levels and
hence cause precipitation. However, it was discovered [19] using aqueous CaCl2 and
K2SO4 mixtures that the BCE process actually has a significant inhibition effect on the
precipitation process, as shown, for example, in Figure 27. In these particle growth
studies turbidity was monitored with time during the precipitation. In this case, with no
added foreign nucleating particles, turbidity values within the BCE solutions were fairly
constant with time over more than 300 min, even though in quiescent solutions, at this
same supersaturation level, significant particle growth was observed much earlier (see
Figure 27).
These results suggest that the high density of rising bubbles might disrupt sub-nu-
clei or molecular clusters in the solution during precipitation and, in addition, the
charged surfaces of the bubbles might have a big perturbation on interacting ions
involved in nucleation and growth. A similar phenomenon was reported by Mullin and
Raven [91,92] who found that an increase in the intensity of agitation does not always
lead to an increase in nucleation, which might be explained by assuming that agitation
can disrupt sub-nuclei or molecular clusters in the solution. Stirred systems are also
complex in that, for example, different types of impellers can have a significant effect
on other process parameters of crystallisation [93] and can lead to the production of
different crystal shapes [94].
International Reviews in Physical Chemistry 179

Figure 27. (Colour online) CaSO4·2H2O precipitation at ionic product of 0.042 M2 in stirring
(cylindrical stir bar of 3.5 cm length and 1 cm diameter) and BCE systems monitored by a tur-
bidity meter at around 25 °C (reproduced from Fan and Pashley [19]).

The growth of CaSO4·2H2O precipitates with time in both BCE and stirring pro-
cesses was monitored using the Malvern zeta-sizer, and an example of typical results is
given in Figure 28, where the ionic product was 0.042 M2 and the precipitation was
nucleated with 23 nm diameter silicon dioxide spheres (at 250 ppm). It can be seen
from these results that the BCE process, compared to simple stirring, inhibited hetero-
geneous nucleation and particle growth of the CaSO4 precipitates. These results also
suggest that the BCE process for precipitation could potentially be used as a method of
controlled particle growth and in the production of fine particles.
Images of CaSO4 precipitates at different times from the BCE and stirring processes
were captured using an atomic force microscope (AFM) and scanning electron micro-
scope (SEM), as shown in Figure 28. The particle sizes produced in the BCE process
from 20 to 40 nm in AFM images and 10 to 20 μm in SEM images, appear to show
that the precipitate first deposited onto the surface of spherical silica, produced slightly
elongated particles (see BCE at 150 min from AFM in Figure 28) and then became
needle-like after a certain time in the BCE process. By comparison, precipitation by
stirring was fast and the precipitates had a ripened appearance (needle-like) within a
shorter time [19].

3.6.3. The application of precipitation inhibition with the BCE


Comparison of the BCE experiments with stirring studies suggests that the BCE pro-
cess, due to the effect of inhibition of crystal nucleation and growth, could be impor-
tant and helpful in particle growth control applications, as well as in the treatment of
heavily contaminated industrial wastewater. Often, the aim of wastewater treatment is
to concentrate the solutions as much as possible prior to disposal, for example, in evap-
oration ponds. Industrial waste vent gases and waste heat or solar heat can be used for
the BCE process with heated inlet gases but without solution boiling, which could be a
simpler and more efficient way for concentrating contaminated waste water. Alternative
methods, such as filtration can be costly because of the need for renewal of membranes
and boiling has high energy costs and quickly becomes inefficient due to solute precipi-
180 M. Shahid et al.

Figure 28. (Colour online) The growth of CaSO4·2H2O particles at an ionic product of 0.042 M2
in BCE and stirring systems, with 23 nm diameter silica spheres at around 25 °C. SEM and
AFM images were taken from samples withdrawn at different times (adapted from Fan and
Pashley [19]).

Figure 29. (Colour online) Estimated bubble column (0.5 m NaCl solution) temperature at 3 atm
pressure using different inlet gas temperatures and the water vapour density at different column
temperatures (reproduced from Francis and Pashley [14]).
International Reviews in Physical Chemistry 181

tation and fouling of the heat-exchange pipes. By comparison, during the BCE process
saturated water vapour can be collected and condensed to access a continuous source
of high quality water [19,21].
Using this non-boiling BCE process, as an example, the calciner flue gases pumped
at typically around 165 °C, into a BCE at 3 atm pressure will maintain the column tem-
perature at around 60 °C [14] (see Figure 29, calculated from Equation (7)), which
would be capable of producing high quality water from 0.5 m NaCl solution at the rate
of 0.13 L m−3 of hot gases.

3.7. Other potential applications


The BCE process, in addition to its applications in waste water concentration, could at
the same time be used to inactivate different types of viruses and enzymes in the
wastewater. The BCE hot air system could also possibly be used for the sterilisation of
dairy and dairy based products and could even be used for the production of chiral
compounds. In addition, the bubble column system could be effective in treating water
based foods, beverages, blood and blood related products and it might also be
employed in specific stages for the treatment of pharmaceutical products. The BCE
could also be used for carbon capture by adding surfactant to produce a continuous
flow of CO2 foam from hot waste industrial gases. This foam could be transported and
then buried under pressure to reduce total volume and foam cell size for long term
storage.

4. Conclusion
This review article examines the theoretical background and the applicability of the
BCE process to several important industrial applications. The design and development
of the BCE are established on two main features (1) higher overall mass transfer coeffi-
cient and (2) efficient heat transfer coefficient, which is a prerequisite for different BCE
applications. It is argued that a thorough and in-depth understanding of the BCE sys-
tem, when used with aqueous solutions, is critically dependent on the physical proper-
ties of water. Further detailed studies would be required to develop large scale
industrial applications of this technique.

Nomenclature
ΔHvap Enthalpy of vaporisation
FB Buoyancy force
Fcp Contact pressure force
Fc Surface tension or capillary force
FD Dynamic force
ΔP Pressure difference between inside and outside of the bubble
γ Surface tension of water
rw Radius of the water molecule
r Radius of air bubble
T Absolute temperature
ΔT Temperature difference between inlet gas and bubble column
solution
ΔT′ Temperature difference between inlet gas and bubble surface
182 M. Shahid et al.

PHB Degree of hydrogen bonding


ΔEHB Energy required for completely breaking and separating water
molecules
ΔWv External work
ΔUv Surface energy
S Cross-sectional area of the column
dia Diameter
U∞ Bubble rising velocity within the column at steady state
rf Flow rate monitored by the flow meter
Trm Column solution temperature at steady state condition
Tf Gas temperature at the flow meter
Pc Gas pressure at the column top
Pf Gas pressure at the flow meter
θa Collision angle
θt Maximum collision angle
tbl Bubble layer thickness
Mw Molar mass of water
qvw ðTe Þ Water vapour density at steady state column temperature
qw Liquid water density
G Gravitational constant
qg Density of air (or gas)
πos Osmotic pressure
Cp Specific heat capacity at constant pressure of air
Cs Specific heat capacity of solution (or water)
i van’t Hoff coefficient
P Saturated water vapour pressure
rc Concentrating rate in BCE process per unit of time
ci and c0i Initial and final molar concentration of salt solution i
δ Hot layer thickness
η The coefficient of viscosity of liquid

Acronyms
BCE Bubble column evaporator
MSF Multistage flash distillation
MED Multi-effect distillation
SWRO Seawater reverse osmosis
VC Vacuum compressor
DLVO Theory presented by Derjaguin, Landau, Verwey and Overbeek
VDW van der Waals
UV Ultraviolet
HFC High faecal coliform numbers
MFC Medium faecal coliform numbers
LFC Low faecal coliform numbers
ANFC Almost no faecal coliform numbers

Non-SI units
min Minute
atm Atmospheric pressure
NTU Nephelometric Turbidity Unit
International Reviews in Physical Chemistry 183

Disclosure statement
No potential conflict of interest was reported by the authors.

Funding
This work was supported by the Australian Research Council [grant number DP120102385].

References

[1] P. Zehner and M. Kraume, Ullmann’s Encyclopedia of Industrial Chemistry (Wiley, 2000).
[2] M. Shahid, X. Xue, C. Fan, B.W. Ninham and R.M. Pashley, J. Phys. Chem. B 119 (25),
8072–8079 (2015).
[3] C. Fan, M. Shahid and R.M. Pashley, J. Solution Chem. 43 (8), 1297–1312 (2014).
[4] M.J. Francis and R.M. Pashley, J. Phys. Chem. B 113 (27), 9311–9315 (2009).
[5] V.S.J. Craig, B.W. Ninham and R.M. Pashley, J. Phys. Chem. 97 (39), 10192–10197 (1993).
[6] A.H. Luedicke, B. Hendrickson and G.M. Pigott, J. Food Sci. 44 (5), 1469–1473 (1979).
[7] C.P. Ribeiro and P.L.C. Lage, Chem. Eng. Technol. 28 (10), 1081–1107 (2005).
[8] Y. Shah, B.G. Kelkar, S. Godbole and W.D. Deckwer, AIChE J. 28 (3), 353–379 (1982).
[9] V.I. Klassen and V.A. Mokrousov, in An Introduction to the Theory of Flotation, edited by
J. Leja and G.W. Poling (Butterworths, London, 1963).
[10] V.S.J. Craig, Curr. Opin. Colloid Interface Sci. 9 (1–2), 178–184 (2004).
[11] S. Marčelja, Curr. Opin. Colloid Interface Sci. 9 (1–2), 165–167 (2004).
[12] R.M. Pashley and V.S.J. Craig, Langmuir 13 (17), 4772–4774 (1997).
[13] C. Fan and R.M. Pashley, J. Solution Chem. 44 (1), 131–145 (2015).
[14] M.J. Francis and R.M. Pashley, Desalin. Water Treat. 12 (1–3), 155–161 (2009).
[15] M. Shahid and R.M. Pashley, Desalination 351, 236–242 (2014).
[16] X. Xue and R.M. Pashley, Desalin. Water Treat. 57 (20)9444–9454 (2016).
[17] M. Shahid, J. Water Process Eng. 8, e1–e6 (2015).
[18] M. Shahid, R.M. Pashley and R.A.F.M. Mohklesur, Desalin. Water Treat. 52, 4444–4452
(2014).
[19] C. Fan and R.M. Pashley, Chem. Eng. Sci. 142, 23–31 (2016).
[20] P.N. Govindan, G.P. Thiel, R.K. McGovern, J.H. Lienhard and M.H. Elsharqawy, Patent
No. US8523985 (3 September 2013).
[21] G.P. Narayan, J.H. Lienhard and I.D.A.J. Desalin, IDA J. Desalin. Water Reuse 4 (3), 24–34
(2012).
[22] M. Schmack, H. Goen and A. Martin, Environ. Technol. 37 (1), 74–85 (2015).
[23] J. Davidson and B. Schuler., Chem. Eng. Res. Des. 75, S105–S115 (1997).
[24] A. Khurana and R. Kumar, Chem. Eng. Sci. 24 (11), 1711–1723 (1969).
[25] S. Ramakrishnan, R. Kumar and N. Kuloor, Chem. Eng. Sci. 24 (4), 731–747 (1969).
[26] S.D. Bari and A.J. Robinson, Exp. Therm. Fluid Sci. 44, 124–137 (2013).
[27] R. Bowers, J. Appl. Chem. 5 (10), 542–548 (1955).
[28] A. Satyanarayan, R. Kumar and N. Kuloor, Chem. Eng. Sci. 24 (4), 749–761 (1969).
[29] R.M. Pashley and M.E. Karaman, in Applied Colloid and Surface Chemistry, edited by
Geoffrey R. Luckhurst and Timothy J. Sluckin (Wiley, Chichester, 2004).
[30] P. Ghosh, Chem. Eng. Res. Des. 82 (7), 849–854 (2004).
[31] V. Craig, B. Ninham and R. Pashley, Nature 364 (6435), 317–319 (1993).
[32] C. Foulk, Ind. Eng. Chem. 21 (9), 815–817 (1929).
[33] A.K. Chesters, Chem. Eng. Res. Des. 69 (A4), 259–270 (1991).
[34] Y. Liao and D. Lucas, Chem. Eng. Sci. 65 (10), 2851–2864 (2010).
[35] W. Howarth, AIChE J. 13 (5), 1007–1013 (1967).
[36] W. Howarth, Chem. Eng. Sci. 19 (1), 33–38 (1964).
184 M. Shahid et al.

[37] F. Lehr, M. Millies and D. Mewes, AIChE J. 48 (11), 2426–2443 (2002).


[38] P. Duineveld, Ph. D. thesis, Twente University, 1994.
[39] L. Doubliez, Int. J. Multiphase Flow 17 (6), 783–803 (1991).
[40] R.G. Horn, L.A. Del Castillo and S. Ohnishi, Adv. Colloid Interface Sci. 168 (1–2), 85–92
(2011).
[41] J.N. Israelachvili, Intermolecular and Surface Forces, 2nd ed. (Academic Press, London,
1992).
[42] J. Israelachvili and R. Pashley, Nature 300, 341–342 (1982).
[43] H.K. Christenson and P.M. Claesson, Adv. Colloid Interface Sci. 91 (3), 391–436 (2001).
[44] Y. Katsir and A. Marmur, Langmuir 30 (46), 13823–13830 (2014).
[45] Y. Katsir and A. Marmur, Sci. Rep. 4 (2014).
[46] J. Garai, Fluid Phase Equilib. 283 (1–2), 89–92 (2009).
[47] Y. Zhang and Z. Xu, Am. Mineral. 80 (7–8), 670–675 (1995).
[48] D.R. Lide and T.J. Bruno, CRC Handbook of Chemistry and Physics, (CRC Press, Boca
Raton, FL, 2012).
[49] C.D. Cappa, J.D. Smith, K.R. Wilson, B.M. Messer, M.K. Gilles, R.C. Cohen and R.J.
Saykally, J. Phys. Chem. B 109 (15), 7046–7052 (2005).
[50] I. Leifer, R.K. Patro and P. Bowyer, J. Atmos. Oceanic Technol. 17 (10), 1392–1402
(2000).
[51] R.M. Pashley, M.J. Francis and M. Rzechowicz, Water 35 (8), 67–71 (2008).
[52] A.A. Kulkarni and J.B. Joshi, Ind. Eng. Chem. Res. 44 (16), 5873–5931 (2005).
[53] R. Clift, J. Grace and M. Weber, Bubbles, drops, and particles (Academic Press, New York,
NY, 1978).
[54] J. Quinn, M. Maldonado, C. Gomez and J. Finch, Miner. Eng. 55, 5–10 (2014).
[55] P. Gonzalez-Tello, F. Camacho, E. Jurado and R. Bailon, Can. J. Chem. Eng. 70 (3),
426–430 (1992).
[56] X. Luo, J. Zhang, K. Tsuchiya and L.-S. Fan, Chem. Eng. Sci. 52 (21–22), 3693–3699
(1997).
[57] K.W.K. Li and A. Schneider, J. Am. Ceram. Soc. 76 (1), 241–244 (1993).
[58] V.G. Levich and S. Technica, in Physicochemical Hydrodynamics, edited by Editor
(Prentice-Hall, Englewood Cliffs, NJ, 1962).
[59] D.Y. Chan, E. Klaseboer and R. Manica, Soft Matter 7 (6), 2235–2264 (2011).
[60] E. Klaseboer, R. Manica, D.Y. Chan and B.C. Khoo, Eng. Anal. Boundary Elem. 35 (3),
489–494 (2011).
[61] E.W. Lemmon, R.T. Jacobsen, S.G. Penoncello and D.G. Friend, J. Phys. Chem. Ref. Data
29 (3), 331–385 (2000).
[62] R.M. Pashley, Patent No. US20090120877 (14 May 2009).
[63] R.G. Lunnon, Proc. Phys. Soc. London 25, 180–191 (1912).
[64] B.M. Gurovich, R.A. Zyuzin and S.M. Mezheritskii, Chem. Pet. Eng. 24 (11), 599–600
(1988). doi:10.1007/bf01154379.
[65] F.J. Massey Jr., J. Am. Stat. Assoc. 46 (253), 68–78 (1951).
[66] B.A. Adams and E.L. Holmes, J. Soc. Chem. Ind. 54, 1–6 (1935).
[67] M.H.I. Dore, Desalination 172 (3), 207–214 (2005).
[68] A.D. Khawaji, I.K. Kutubkhanah and J.-M. Wie, Desalination 221 (1–3), 47–69 (2008).
[69] G. Fiorenza, V. Sharma and G. Braccio, Energy Convers. Manage. 44 (14), 2217–2240
(2003).
[70] M. Al-Sahali and H. Ettouney, Desalination 214 (1–3), 227–240 (2007).
[71] I.C. Karagiannis and P.G. Soldatos, Desalination 223 (1–3), 448–456 (2008).
[72] M. Shahid and R.M. Pashley, in Aqua Incognita: Why Ice Floats on Water and Galileo
400 Years On, edited by P. Lo Nostro and B.W. Ninham (Connor Court, Ballarat, 2014), pp.
350–366.
[73] M.P. Doyle, L.R. Beuchat and T.J. Montville, Food Microbiology: Fundamentals and
Frontiers, 2nd ed. (ASM Press, Washington, DC, 2001), p. 872.
International Reviews in Physical Chemistry 185

[74] E.E. Geldreich, H.F. Clark, B.H. Clifton and L.C. Best, J. Am. Water Works Assoc. 57 (2),
208–214 (1965).
[75] UNIDO, OPEC Rev. 8 (4), 409–433 (1984). doi:10.1111/j.1468-0076.1984.tb00287.x.
[76] R.M. Pashley, X. Xue, C. Fan, B.W. Ninham and M. Shahid, Patent No. AU2015901956
(27 May 2015).
[77] J.R. McCutcheon, R.L. McGinnis and M. Elimelech, Desalination 174 (1), 1–11 (2005).
[78] N.P.G.N. Chandrasekara and R.M. Pashley, Desalination 357, 131–139 (2015).
[79] M. Okubo and T.Mori, Makromolekulare Chemie. Macromolecular Symposia (Wiley Online
Library, Basel,, 1990), Vol. 31, pp. 143–156.
[80] G. Fulks, G.B. Fisher, K. Rahmoeller, M.-C. Wu, E. D’Herde and J. Tan, SAE Technical
Paper (2009). doi:10.4271/2009-01-0907.
[81] G.W. Gokel, in Dean’s Handbook of Organic Chemistry, edited by G.W. Gokel
(McGraw-Hill, New York, 2004).
[82] I. Kolthoff and I. Miller, JACS 73 (7), 3055–3059 (1951).
[83] J.C. Cowan and D.J. Weintritt, in Water-formed Scale Deposits, edited by Editor (Gulf
Publishing Company, Book Division, Houston, 1976).
[84] S. He, J.E. Oddo and M.B. Tomson, J. Colloid Interface Sci. 162 (2), 297–303 (1994).
[85] S. He, J.E. Oddo and M.B. Tomson, J. Colloid Interface Sci. 163 (2), 372–378 (1994).
[86] P.G. Klepetsanis and P.G. Koutsoukos, J. Colloid Interface Sci. 143 (2), 299–308 (1991).
[87] B. Smith and F. Sweett, J. Colloid Interface Sci. 37 (3), 612–618 (1971).
[88] C. Hazra, S. Bari, D. Kundu, A. Chaudhari, S. Mishra and A. Chatterjee, Ultrason.
Sonochem. 21 (3), 1117–1131 (2014).
[89] Y.-W. Wang, Y.-Y. Kim, H.K. Christenson and F.C. Meldrum, Chem. Commun. 48 (4),
504–506 (2012).
[90] H. Tiemann, I. Sötje, G. Jarms, C. Paulmann, M. Epple and B. Hasse, J. Chem. Soc., Dalton
Trans. 7, 1266–1268 (2002).
[91] J.W. Mullin and K.D. Raven, Nature 195, 35–38 (1962).
[92] J.W. Mullin and K.D. Raven, Nature 190 (4772), 251–251 (1961).
[93] W. Beckmann, in Crystallization: Basic Concepts and Industrial Applications, edited by W.
Beckmann (Wiley, 2013).
[94] I. Mukhopadhyay, V.P. Mohandas, G.R. Desale, A. Chaudhary and P.K. Ghosh, Ind. Eng.
Chem. Res. 49 (23), 12197–12203 (2010).
[95] E.C.W. Clarke and D.N. Glew, J. Phys. Chem. Ref. Data 14, 489–610 (1985).
[96] D.E. Garrett, in Handbook of Lithium and Natural Calcium Chloride, edited by Editor
(Academic Press, Ojai, 2004).
[97] K.R. Patil, A.D. Tripathi, G. Pathak and S.S. Katti, J. Chem. Eng. Data 36 (2), 225–230
(1991).

S-ar putea să vă placă și