Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Desalination 420 (2017) 258–272

Contents lists available at ScienceDirect

Desalination
journal homepage: www.elsevier.com/locate/desal

CFD prediction of the flashing processes in a MSF desalination chamber MARK



T.H. Nigim , J.A. Eaton
Mechanical Engineering, National University of Ireland Galway, Ireland

A R T I C L E I N F O A B S T R A C T

Keywords: Vapourisation inside a desalination flashing chamber arises from the energetic evaporation with self-boiling of
Flashing process liquid brine flow due to a reduction of pressure, and is a complex, multiphase, thermo-fluid phenomenon. A
MSF chamber computational field model for the flashing process inside a flashing chamber is developed around a two-phase
Evaporation zone VOF formulation. Two different phase-change mechanisms are allowed for, based on the saturation temperature
CFD prediction
and on the vapour pressure, respectively, that enable the model to compute the phase change regions, also the
Multiphase
level and the shape of the free surface. The model is applied to solve for steady multiphase flow inside a flashing
chamber without a baffle, and thus to predict the chamber flow details and behaviour. Thermo-fluid perfor-
mance results are presented in terms of flow patterns and distributions of hydrostatic pressure, temperature,
vapour volume fraction and mass transfer inside the flashing chamber. The average outlet temperature and
average vapour temperature agree well with the values of a real MSF plant, as does the estimated vapour
production rate. Two main mechanisms of phase change produced by the flashing process are captured in the
simulation. Vapour bubbles are formed at the entrance to the flashing chamber, and bubble production reduces
along its length. At the free surface of the liquid there is also a phase change and mass transfer. Computational
field modelling is valuable for a better understanding of the combined heat transfer, mass transfer and the fluid
dynamics during the flashing flow evaporation processes. Results can be used to estimate MSF design factors
such as non-equilibrium temperature difference and flashing efficiency.

1. Introduction Fig. 1 shows a schematic diagram of a horizontal flow MSF flashing


chamber, together with a general configuration of the evaporation
The multi-stage flash (MSF) desalination system was among the zone. The evaporation zone is fed with brine through a submerged inlet
earliest widely-used approaches to desalination [1], and continues as orifice, and comprises an open flow channel fitted with a flow baffle
one of the most common and simple methods in current use [2]. Cap- leading to a submerged outlet: the outlet forms the inlet orifice to the
able of producing fresh water of a very high purity, it is also suitable for next stage. The stage consists of the following basic elements: inlet
the largest-scale production. Indeed, since MSF relies on thermal input, orifice, brine pool, flow baffle, splash baffle, outlet orifice, demister,
it is preferably to be built as a cogeneration plant for concurrent pro- condenser/preheater tubes, distillate tray, venting line and partition
duction both of power and of fresh water [3]. Notwithstanding the walls. In a large-scale desalination plant, a flashing stage typically [6]
introduction of many alternative desalination processes, MSF remains has width, height, and length dimensions of approximately
very important: in 2014 it accounted for about 22% of global desali- 18 m × 4 m × 3 m, respectively.
nation plant production [4,5]. Thermodynamically, the flashing chamber has the highest exergy in
The fundamental principles in the multi-stage flashing flow process the MSF system [7], and the evaporation zone may have the highest
consist first of producing water vapour from the brine, then of con- exergy within the flashing chamber [8]. In this respect, the evaporation
densing the vapour to yield distilled water. The key mechanism is the zone is thus considered as the most important part of the stage [1], and
changing of phase of the brine, which is attained by causing the brine is the focus of this study.
pressure to fall below its vapour pressure: once the brine becomes The flashing chamber maybe treated as an open system, mainly
sufficiently superheated, and in the presence of nuclei, vapour bubbles because the boundaries of the system are permeable both to energy and
are produced as a consequence of phase change of the water. The re- to mass. In general, the flow field in the evaporation zone may be
duction in pressure is arranged stepwise, in a continuous flow process, classified as multiphase, complex, turbulent and unsteady, featuring a
to occur in successive series of flashing chambers known as stages. submerged orifice flow, a wall jet, flow recirculation and an


Corresponding author.
E-mail address: t.nigim1@nuigalway.ie (T.H. Nigim).

http://dx.doi.org/10.1016/j.desal.2017.06.026
Received 18 July 2016; Received in revised form 14 June 2017; Accepted 25 June 2017
0011-9164/ © 2017 Elsevier B.V. All rights reserved.
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

Nomenclature vv velocity vector of the vapour phase [m/s]


vy component of velocity in y-direction [m/s]
coeff rate factor of vapourisation or condensation (inverse of a w width of flashing chamber orifice [m]
relaxation time) in Lee Wen Ho model (Eq. (3)) [s− 1] W width of flashing chamber [m]
d vapour bubble diameter [m] x horizontal coordinate [m]
Fcol collapse coefficient [−] y vertical coordinate [m]
Fvap vapourisation coefficient [−] y+ wall-normal distance ≈ y ρτw μ [−]
g gravitational acceleration [m/s2]
h height of inlet orifice to flashing chamber [m] Greek
H height of flashing chamber [m]
L length of flashing chamber [m] α volume fraction [−]
LH latent heat [J/kg] αl liquid phase fraction [−]
ṁ l → v mass transfer rate from liquid to vapour phase [kg m3/s] αnuc nucleation site volume fraction [−]
ṁ v → l mass transfer rate from vapour phase to liquid [kg m3/s] αv vapour phase fraction [−]
M molecular weight [kg/kmol] β parameter in Lee Wen Ho phase-change model (Eq. (5))
n bubble density number [−] [m/s]
n unit vector, outward-normal to S [−] ε dissipation rate of turbulence kinetic energy [m2/s3]
NETD non-equilibrium temperature difference [K] k kinetic energy of turbulence [m2/s2]
ps′ gauge pressure [Pa] μ viscosity [Pa s]
p absolute pressure [Pa] ρl density of liquid [kg/m3]
pop flashing chamber operating pressure ρv density of vapour [kg/m3]
pv vapour pressure [Pa] σ surface tension [N/m]
q heat flux [W/m2] τw wall shear stress [Pa]
rc critical radius of the bubble [m] Ω control volume [m3]
R universal gas constant [J/kmol K]
R′ rate of mass change of a single bubble [kg/s] Subscripts
Rb bubble radius [m]
Rc mass transfer source term, bubble collapse (Eq. (6)) b bubble
[kg m3/s] i number of the stage (flashing chamber) in MSF plant
Re mass transfer source term, bubble growth (Eq. (6)) in inlet
[kg m3/s] l liquid phase
S control surface [m2] nuc nucleation
t time [s] out outlet
T temperature [K] op flashing chamber operating pressure
Tin inlet temperature [K] v vapour
Tl temperature of the liquid phase [K]
Tout outlet temperature [K] Abbreviations
Tsat saturation temperature [K]
Tv temperature of the vapour phase [K] CFD computational fluid dynamics
vb velocity of control surface [m/s] MSF multi-stage flash
vx component of velocity in x-direction [m/s] VOF volume-of-fluid

unconstrained, open surface. Interactions exist between the liquid and effects may be expected to dominate. Moreover, the degree of sub-
vapour phases involving mass, momentum and energy exchanges re- mergence of the inlet, characterised by the ratio of brine level at inlet/
lated to vapourisation and condensation mass transfer, mean flow inlet orifice height, say, can also strongly influence the flow field, as
gradients and turbulence, buoyancy and heat transfer. will the geometric aspect ratio of the chamber.
Here, for simplicity, we consider two-dimensional flow in a flashing Within the brine flow layer, under a process of heterogeneous
chamber without any flow baffles. According to the phase type and to flashing vapour bubbles emerge on nuclei of dissolved gas or of solid
the fluid behaviour, the chamber space may be regarded as divided into particles entrained in the brine. Vapour formation within the brine
a number of layers and regions. These are indicated in Fig. 2 (adapted depends on the local pressure and temperature: brine pressure pL de-
from Lior [9]). In the horizontal sense, there are two main regions: A – pends on the local depth below the free surface. The local difference
the bubble nucleation region, near the inlet, which includes the sub- between the vapour pressure in a bubble and the brine pressure, the
merged jet and recirculating flow; and B - the channel flow, further superheat term (pv − pl), is the key factor in determining whether the
downstream, which is predominantly unidirectional. In the vertical bubble will nucleate and grow, or collapse and disappear.
direction, the chamber space may be considered as comprising three
layers: I – the brine flow layer, near the base; II – the free surface; and
1.1. Purpose of the present work
III - the vapour layer.
The flow may be generally classified as an open channel flow driven
Many studies [11–40] — experimental, analytical, numerical si-
by a submerged jet [10]. The submerged orifice functions as a sluice
mulation — have been undertaken with a view to understanding the
gate, controlling the flow rate and brine level. Such flows are char-
mechanisms behind the evaporation zone and the flashing phenomena,
acterised by Froude number and Reynolds number. Here the Froude
as well as to propose methods for improving the efficiency and per-
number, based on brine depth and mean inlet speed, lies around the
formance of the evaporation zone.
critical value, which may result in a wide variety of surface waves and
Our aim is to produce and validate a multiphase computational field
unsteady flow patterns: under these circumstances Froude number
approach to predict the flow, heat and mass transfer in the evaporation

259
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

zone of an MSF stage. Once fully validated, the method is applicable to interfacial stresses as well as surface tension effects were ignored in the
assisting in the design procedure for MSF systems. calculations. In practice, the flow is unsteady, but even in time-average
This paper is organised as follows: we next review (§1.2) some of the [43] the free surface can deviate significantly from being flat. No
previous computational studies for predicting the evaporation zone. consideration was given to the vapour layer, or to interactions between
Our computational model for the flashing chamber is presented in §2, the brine and vapour layers. Since the model did not have the ability to
together with a validation case based on an existing stage. An outline of predict either the brine level or the free surface shape, the imposition of
the mathematical modelling and computational procedure is also pro- unrealistic free surface conditions will have adversely affected the
vided. In §3 we show the field prediction results of the thermo-fluid predicted thermo-fluid behaviour, and the pressure distribution and
performance (hydrostatic pressure, flow patterns, thermal performance, flow pattern in particular.
vapour volume fraction and the mass transfer) inside the flashing In subsequent work, Seul et al. [44] simplified and refined their
chamber. Results are discussed in §4, and our conclusions are presented bubble dynamic model, void fraction correlation and boundary condi-
in §5. tions, and applied their improved model to investigate the effects of the
liquid level on flow behaviour inside a multi-stage flash evaporator.
They confirmed that bubble motions are very much dependent on the
1.2. Review of computational predictions for the evaporation zone inlet velocity profile, and showed that the evaporation performance is
improved by employing lower liquid levels. Nonetheless, they retained
Seul et al. [41] presented a computational model for simulating the the free surface boundary condition as horizontal with zero shear.
thermal-hydrodynamic behaviour of the horizontal stream inside a Miyatake et al. [43] employed a two-dimensional, incompressible,
multi-stage flash evaporator without a flow baffle. The PSI-CELL steady, single-phase, turbulent, isothermal, stream function-vorticity
method of Crowe et al. [42] for disperse particle flows was adapted and numerical model in an attempt to reach a quantitative understanding of
used to solve the two-dimensional, incompressible, steady, two-phase, the interaction between fluid mechanics and flash evaporation in a
turbulent flow conservation equations of mass, momentum and energy flashing chamber with and without a flow baffle. Although a simple
in conjunction with ancillary equations for motion and growth of va- one-equation, mixing-length turbulence model was used, the flow field
pour bubbles. The liquid (continuous) phase was treated using the contained a number of distinctly different flow zones. For accurate flow
Eulerian field approach, whereas the vapour (disperse) phase was computations, each flow zone has a characteristic (mixing) length scale
computed with a Lagrangian model. Bubble nuclei entered the domain which needs to be calculated. However, the number, type and shape of
at the same velocity as the surrounding liquid, with the initial size and flow zone depends on the configuration of the flashing chamber, and
distribution of bubble nuclei on the inlet boundary specified as para- are not know a priori; thus determination of such length scales is
meters. Cell-averaged bubble trajectories and cell-averaged bubble size cumbersome, and requires an iterative process starting, for instance,
were computed as part of the solution. The method allowed for inter- with a laminar field solution. The curved free surface shape was ob-
actions between the liquid phase and the vapour bubbles, in terms of tained from laboratory experiments, and prescribed as a zero shear
mass transfer (by evaporation), momentum transfer via interfacial stress boundary condition.
forces and energy transfer by vapourisation. Calculated results of the Miyatake et al. [45] studied the relationship between the flashing
flow patterns, in the form of liquid streamlines, temperature distribu- flow pattern and thermal non-equilibrium in a flashing chamber with
tions and bubble trajectories, were presented. Seul et al. [41] employed flow baffles of different locations and heights. They post-processed the
a standard k - ε turbulence model equation for the liquid phase only. velocity and pressure distribution results from their previous study
Only the brine layer was treated, and the free surface level and profile [43], and took the computed streamlines as paths along which to solve
was prescribed. The temperature boundary condition at the free surface for the temperature drop between inlet and exit, for which they em-
was determined from the convective heat transfer rate with surface ployed an empirical correlation for the temperature decrease rate of a
evaporation, based on kinetic theory. An unrealistic flat (horizontal) flashing liquid. By averaging over several streamlines they computed
free surface profile was specified as a boundary condition, and

Fig. 1. Schematic of a flashing chamber shape and evaporation zone elements.

260
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

Fig. 2. Flashing chamber without flow baffle – regions and zones (after Lior [9]).

the average temperature of the liquid at the exit. They found that the 2. Current computational model for the flashing chamber
non-equilibrium temperature difference, NETD = Tout − Tv, is reduced
by increasing the brine temperature, also when the flow rate of the A numerical procedure for flashing flow is developed in this present
brine is increased. work and is applied to predict the multiphase flow field in the eva-
Jin and Low [35] investigated the single-phase flow patterns in a poration zone of an idealised, horizontal flow flashing chamber. Our
model flash evaporation chamber both experimentally, by making multiphase model is based on the FLUENT volume-of-fluid (VOF) code
particle image velocimetry measurements, and numerically, by per- implementation, and encapsulates two interaction mechanisms for
forming complementary two-dimensional, incompressible, steady, phase change — based on saturation temperature and vapour pressure,
single-phase, isothermal, turbulent flow field computations. The mea- respectively — that arise during the flashing process.
sured free surface profiles and inlet gate velocity distributions were In addition to including phase change due to both pressure and
used as boundary conditions, and a standard k – ε turbulent flow model thermal effects, our method is novel in introducing and demonstrating the
was employed. Using streamline plots, for a simple horizontal chamber facility to predict the profile as well as the elevation of the free surface,
without a baffle, they showed that the shape of the free surface as well both of which affect the pressure distribution in the brine, and thus are
as details of the recirculating flow pattern depend strongly on the water central to stage performance.
level, flow rate, inlet gate height and inlet velocity distribution. We note that the method is capable of predicting the fully three-
Mansour et al. [46] used the CFX code to predict two-dimensional, dimensional, time-dependent behaviour of the evaporation zone, which
incompressible, steady, two-phase, adiabatic, turbulent flow of the exhibits oscillating, unsteady flow phenomena characteristic of open
flashing process inside a flashing chamber with different numbers and channel flows. However, for such calculations the computational time
locations of triangular baffles. They employed a two-fluid (liquid water and cost is very high: moreover, assimilating and presenting the copious
and water vapour) Euler–Euler, flow formulation in conjunction with a results of a transient analysis is time-consuming and challenging;
k – ε turbulence model. The flow was treated as two interpenetrating combined with an absence of measurement data for validation, it ap-
continua, with individual conservation laws for mass and momentum pears premature to allocate limited resources to such a simulation. For
for each phase. The space occupied by each phase within a computa- pragmatic reasons, therefore, we confine our attention here to running
tional cell is defined by its local volume fraction, and is part of the flow a steady-state, two-dimensional model of the evaporation zone and to
field solution. Measures of steam mass production and pressure drop of analysing the results of these predictions.
the water were used for comparison purposes between different con- This model is applied to the evaporation zone of an existing MSF de-
figurations of baffle, and the computations indicated that the location salination system for which some data are available. Processing of the
of baffles strongly affects the thermo-fluid performance. Results results of the flow field computation enables a visualisation of the flashing
showing contours of vapour volume fraction and of temperature, process, and phase change regions can be seen over the computational
pressure and velocity for the water component were presented. How- domain. The fluid process parameters of hydrostatic pressure and particle
ever, conventional flow patterns, in the form of streamlines or path- paths (streamlines in steady flow) of flashing flow inside the evaporation
lines, which are valuable is assimilating and assessing field results, were zone are provided and presented. The thermodynamic process parameters
not provided. The bubble nucleation region did not appear in the re- of temperature distribution, vapour volume fraction, heat transfer rate,
sults. It is unclear what boundary conditions were employed for each of and mass transfer rate inside the evaporation zone are also shown.
the two fluids, and the physics and mechanism for phase change, if any,
were not explained.
As already observed, flashing chamber type flows are very complex 2.1. Mathematical modelling
and difficult to predict. From this review of previous computational
work it is evident that better models are needed that are capable of The choice of multiphase model is determined by the application,
treating a number of significant aspects not hitherto allowed for. The and by the flow phenomena of the different phases that it is desired to
local pressure at a point in the brine layer is influenced by its sub- capture. In this application we simplify the fluids by assuming the
mergence below the free surface; the local pressure affects the vapour presence of two fluid phases only: pure water (liquid) and water va-
saturation pressure, and hence the local saturation temperature. In pour. For the present, therefore, real fluid aspects such brine/seawater
particular, therefore, it is desirable to have an approach that enables the mixture properties, brine concentration, and the presence of entrained
shape, elevation and motion of the free surface to be predicted, rather solids and non-condensable gases are ignored. The concentration of
than to have to specify it empirically, or to impose it unrealistically as a vapour varies widely in the evaporation zone of the flashing chamber
flat, horizontal boundary. Prediction of phase change and mass transfer (Fig. 2). We anticipate a dilute, disperse flow of vapour bubbles in the
arising both from thermal effects (i.e. temperature difference) and from bubble nucleation region near the brine inlet, and a very high vapour
mechanical effects (i.e. pressure difference) is also desirable. We next concentration from the free surface upwards. Apart from predicting the
present a new computational model which aims to address these as- flow fields in the brine and vapour layers, we wish to capture the lo-
pects. cation of the interface between the layers. The combination of a sub-
merged sluice, open-channel flow interacting with recirculation and

261
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

phase change may be strongly unsteady, so the model chosen should be solution. To reach this, the space conservation law (SCL) and boundary
capable of handling these interactions in a robust manner. conditions at the free surface must be used. The SCL is used to apply the
Mathematical modelling of two-phase flashing chamber flow has mass conservation equation in the limit of zero fluid velocity. The fol-
been undertaken by an Euler-Euler, two-fluid approach [46], and by an lowing expression is used to describe the conservation of space when
Euler-Lagrange approach [41,44]. Models for these and several other the control volume shape and/or position changes with respect to time.
multiphase formulations are now available in commercial codes. We
d
have selected an Euler-Euler, one-fluid approach, the Volume of Fluid
dt
∫Ω dΩ − ∫S v b ∙ n dS = 0
(1)
(VOF) method [48–64] as implemented in the FLUENT 14.5 finite vo-
lume code [48,49], which is the most appropriate for the present re- where t is time, Ω represents the control volume, S is the control surface
quirement: in particular, it enables the prediction of free surface flows with unit outward normal vector n , and v b is the control surface ve-
of arbitrary interface shape. locity vector.
We now outline the VOF approach used, together with the thermal In the VOF model the fluids share a single set of momentum equa-
and mechanical models used to predict the flashing processes inside the tions, so that the resulting velocity field is shared among the phases.
evaporation zone. The energy equation is also shared among the phases, and energy and
temperature are treated as mass-averaged variables. Likewise, arising
from solution of a single set of turbulent transport equations, the tur-
2.1.1. Volume of fluid (VOF) multiphase model
bulence variables are shared. The volume fraction of each of the fluids,
In the general Euler-Euler multiphase approach the different phases
in each computational cell, is computed throughout the domain and is
(which may include bubbles and particles as well as multiple fluid
used to identify any interfaces that emerge as part of the solution.
phases) are treated mathematically as interpenetrating continua on a
The results are computed in terms of field properties of the mixture,
fixed Eulerian mesh. The concept of a phasic volume fraction is applied,
together with the phase (or void) fraction field. This means that the VOF
where the volume fractions of the different phases are assumed to be
method does not provide separate information about each individual
continuous functions of space and time, and the sum of the volume
phase — as does an Eulerian model that solves individual momentum and
fractions is unity in each cell. For each phase conservation equations
continuity equations for every phase — but rather of the shared properties
are derived, with a similar equation structure for all phases.
of the (single-fluid) mixture, which may be regarded as its main limitation.
Computing the free surface of flow is an extremely difficult class of
flows involving moving boundaries [49]. The VOF interface tracking
technique is a well-established [50,52,57,60] and validated [51,52,59] 2.1.2. Phase change models
approach, designed for two (or more) immiscible fluids where it is The phase change models based on local thermal (saturation tem-
desired to locate the position of the interfaces between the fluids. perature) and mechanical (vapour pressure) effects, respectively, which
Fig. 3 illustrates a group of rectangular cells in the vicinity of an are used for building the flashing chamber simulation, are presented in
interface. The liquid region is shaded. For the flashing simulation, the this section.
liquid phase fraction is presented as αl while the vapour phase is re-
presented as αv in each cell throughout the domain. Three cell types can 2.1.2.1. Phase change based on thermal effect
be identified: empty cells (where vapour phase αv = 0), cells full of The mechanism of interphase mass transfer between the liquid and
vapour phase (αv = 1), and cells containing the interface between va- the vapour phases is controlled by the vapour mass transfer equation
pour and liquid phase (0 < αv < 1). Based on the local value of αv and [48,63].
αl, appropriate properties and variables will be assigned to each control ∂
volume within the domain. (α v ρv ) + ⋅(α v ρv v v ) = ṁ l → v − ṁ v → l
∂t (2)
At the initial time of the simulation, the position of the boundary is
known. Its location at later times are to be determined as part of the where subscript υ = vapour phase, α = volume fraction, ρv = vapour

Fig. 3. Volume fraction on a discrete mesh.

262
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

density, v v = vapour phase velocity vector, ṁ l → v = mass transfer rate liquid. It is assumed that there are plenty of nuclei present for inception
from liquid to vapour phase, ṁ v → l = mass transfer rate from vapour of cavitation, and the primary focus is on accounting for bubble growth
phase to liquid. and collapse.
The Lee Wen Ho model [64] is a mechanistic model representing The Zwart et al. model [65], which is derived from the generalised
phase change from liquid to vapour phase, as well as from vapour to Rayleigh-Plesset equation [66], is based on the assumption that bubbles
liquid phase, with a physical basis. The mechanism of the vapourisa- in the system are of equal size. While the generalised Rayleigh-Plesset
tion-condensation model is based on checking the temperature of the equation includes the influences of surface tension, liquid viscosity and
liquid phase. Since evaporation is a surface phenomenon, mass transfer inertia on bubble growth, these terms are absent from Zwart et al.
of the molecules takes place at a liquid-vapour interface. When mole- model. In the present application, therefore, bubble growth is assumed
cules have enough kinetic energy, they can escape from the free surface to be dominated by thermal effects.
as a vapour phase. This kinetic energy is a function of the saturation In the Zwart et al. model the bubble density numbers (n) and the
temperature of the liquid. At higher temperatures, molecular kinetic mass change rate of a single bubble are used for calculating the total
energy is greater. interphase mass transfer rate per unit volume (R′):
The local saturation temperature corresponding to the local pressure
DRb
of the system is considered as an indicator for this model. In other R′ = n ⎛4πRb2ρv ⎞
⎝ Dt ⎠ (7)
words, the vapourisation process takes place when the local tempera-
ture of the liquid phase is higher than the local saturation temperature, where Rb is the bubble radius. To determine whether the bubble will
while the condensation process takes place when the temperature of the nucleate and grow, or collapse and disappear, an indicator formula is
vapour phase is higher than the saturation temperature. Thus required. A critical radius for the bubble is the key to creating such an
indicator. The following indicator is used in the model for this purpose:
if Tl > Tsat (vapourisation), then
If the local absolute pressure p ≤ pv, then
(Tl − Tsat )
ṁ l → v = (coeff ) αl ρl
Tsat (3) 3 αnuc (1 − α v ) ρv 2 Pv − P
R e = Fvap
Rb 3 ρl (8)
If Tv > Tsat (condensation), then
(Tv − Tsat )
ṁ v → l = (coeff ) α v ρv If p ≥ pv, then
Tsat (4)
3α v ρv 2 P − Pv
R c = Fcol
Rb 3 ρl (9)
where the coefficient, coeff, must be adjusted carefully in order to −6
Here the generalised bubble radius Rb is set at 10 m, the va-
achieve the best description of performance. It also should be inter-
pourisation coefficient Fvap is set to 50, and the condensation coefficient
preted as a relaxation time. The coefficient factor can be described in
Fcol is set to 0.01. By setting a volume fraction value for nucleation sites
the following expression
as input, in conjunction with the generalised bubble radius the bubble
6 M ρl ⎞ density number is determined. Here, for initial computations, the vo-
coeff = β LH ⎛⎜ ⎟
lume fraction of nucleation sites is specified as 0.0005.
d 2πRTsat ⎝ l ρv ⎠
ρ − (5)

where β is the accommodation coefficient, which defines the proportion 2.2. Validation case study [46,47]
of vapour molecules being adsorbed at the liquid side of the interfacial
surface: β is a physical characteristic of the gas, and has a value of 1.0 The present work is built and validated based on modelling the
near equilibrium conditions; d = vapour bubble diameter, evaporation zone in one stage of an existing MSF plant. The data used
M = molecular weight, R = universal gas constant, LH = latent heat, are for the Sidi Krir Plant at Alexandria, Egypt. The geometry used to
υ = vapour phase, α = vapour volume fraction, ρl = liquid density. define the computational domain and the operating conditions are
The vapourisation-condensation flux is based on the Hertz Knudsen taken from published papers [46,47]. The MSF plant has 20 stages, with
kinetic theory formula for a flat interface. In calculating the interfacial a total production of desalinated water of 5000 m3/day.
area density it is assumed that the flow regime is disperse, also that all The shape and dimensions of the computational domain can be seen
vapour bubbles are spherical with the same diameter. in Fig. 4. Here we simulate the evaporation zone in the first flashing
Here, to enable initial computations in the absence of accurate data chamber of this plant. Operating conditions of the first flashing
for the diameter of the vapour bubbles, and the value of coeff is set to chamber are:
0.1, thus obviating the need to define the diameter.
1. Average inlet temperature of brine 383.15 K
2.1.2.2. Phase change based on mechanical effect 2. Average outlet temperature of brine 379 K
The mass transfer mechanism between liquid-vapour phases are 3. Mass flow rate of the brine 1847 ton/h
based on the vapour transport equation [48]: 4. Flashed vapour 12.5 ton/h
5. Flashing chamber operating pressure 1.023 bar

(α v ρv ) +  . (α v ρv v v ) = R e − R c
∂t (6)
The mass flow rate and inlet temperature of the brine are used to
where v = vapour phase, αv = vapour volume fraction, ρv = vapour define inputs to the model, applied as boundary conditions, while for
density, v v = vapour phase velocity vector, Re = mass transfer source validation the vapour production and temperature field are computed
term connected to the growth of the vapour bubbles, Rc = mass transfer and compared with measured values [46].
source term connected to the collapse of the vapour bubbles. In the
FLUENT cavitation evaporation-condensation model these mass- 2.3. Computational case setup
transfer source terms are modelled based on the Rayleigh-Plesset
equation, which describes the growth of a single vapour bubble in a In this section the simulation geometry with boundary conditions

263
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

Fig. 4. Computational domain: evaporation zone dimensions and boundary conditions


(source data [46]).

Fig. 7. Hydrostatic pressure distribution along horizontal traverses at y = 0.03 m,


0.17 m, free surface line, 0.3 m.

and assumptions are presented. To predict the thermofluid flow field


inside the flashing chamber, we treat the flow as incompressible, two-
dimensional, steady, adiabatic, two-phase (water liquid and water va-
pour) and turbulent.

2.3.1. Computational domain and mesh


The dimensions of the flashing chamber are typical of an existing
stage of a MSF desalination plant (Fig. 4), with a streamwise length of
1.25 m and an inlet gate height of 60 mm. The computational domain is
bounded by walls on three sides, with one inlet and one outlet each for
the brine. Note that the upper (‘open’) boundary for vapour is situated
within the vapour zone, 0.35 m above the floor of the chamber: the
location of the free surface, which is unknown ab initio, will be de-
Fig. 5. Hydrostatic (gauge) pressure field.
termined by the operating conditions, and is computed as part of the
solution.
Quadrilateral rectangular shape elements, with six inflation layers
near the walls, are used for mesh generation. The total number of ele-
ments is 201,358.
Smooth walls are assumed for these computations. In this config-
uration we anticipate a wall jet developing along the floor of the
chamber, and wall regions experiencing flow separation, reattachment
and impingement. Thus, since much of the flow may be wall-domi-
nated, near the solid surfaces the mesh is refined and extended into the
viscous sublayer. The mesh is generated such that along the walls the
estimated non-dimensional wall-normal distance to the centre of the
first cell, y+, is < 5: this enables enhanced wall functions [48] to be
used along with the k – ε turbulence model to better resolve the flows in
these near-wall regions.

2.3.2. Boundary conditions


The boundary conditions applied are shown in Fig. 4. A uniform
inlet velocity profile of 1.5 m/s at 383.15 K is set, with a uniform va-
pour fraction of zero and standard default values for the turbulent
parameters. The base and vertical walls have no-slip and are adiabatic.
The brine outlet and open boundary are set as uniform pressure
boundaries with uniform vapour fractions of zero and one, respectively.
Fig. 6. Hydrostatic pressure distribution along vertical traverses at x = 0.31 m, 0.62 m, For this low-speed, incompressible flow it is appropriate to employ a
0.93 m. pressure-based solution algorithm. In this algorithm the pressure is
derived from iterative solution of the continuity and momentum
equations, and is computed as a gauge-pressure field. Thus all boundary
condition pressures (Fig. 4) are specified to FLUENT as gauge pressures

264
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

Fig. 8. Particle paths (streamlines).

Fig. 9. Mixture speed along vertical traverses at x = 0.31 m, 0.62 m, 0.93 m. Fig. 10. Mixture speed along horizontal traverses at y = 0.03 m, 0.17 m, free surface line,
0.3 m.

(ps′). To enable absolute pressures to be produced a reference pressure


is set at a reference point in the domain. The absolute pressure is then We note that when gravitational acceleration is active in FLUENT, as in
calculated as the sum of the gauge pressure and the reference pressure. the present computations, the hydrostatic component of pressure is
Here the reference pressure is set, on the top of the vapour layer at a included in the gauge pressure field. Hereinafter this gauge pressure
point on the open boundary, to the saturation pressure of the flashing combination will be referred to as the hydrostatic pressure.
chamber (pop = 1.023 bar, corresponding to the saturation temperature The values of fluid properties are estimated from [67,68].
of 373.15 K). The open boundary gauge pressure is therefore set to zero.

265
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

2.4. Numerical schemes and convergence criteria

An implicit pressure-based solver is used in the VOF model to obtain


a steady-state solution. While in reality the flow will be time-depen-
dent, for the present, however, to simplify the analysis and processing
of results we confine consideration to the imposition of an artificial
steady state. To resolve the pressure-velocity coupling the FLUENT
implementation of the PISO (Pressure Implicit with Splitting of
Operators) algorithm is employed, and both the neighbour and skew-
ness corrections are used to improve the efficiency of PISO calculations
[62,69]. FLUENT uses a collocated scheme in which all variables, in-
cluding velocity and pressure, are stored at cell centres. Interpolation is
therefore required to compute the corresponding values of variables
and of their gradients at cell face locations. A least-squares cell-based
interpolation method is used for computing the gradients. The PRESTO
Fig. 11. Temperature field. scheme is employed for pressure interpolation. For improved accuracy,
second-order upwind differencing is applied for the spatial discretisa-
tion of the convective terms in the momentum, volume fraction, tur-
bulence variables and energy equations. During each iteration, for each
cell the mass-transfer and energy interaction between the two phases
are calculated, then returned as source terms to the particular transport
equations.
The convergence criteria are as follows: for each of the x-velocity, y-
velocity, k, ε and volume-fraction-vapour equations it is 10− 5, for the
continuity equation it is 5 × 10− 4 and for the energy equation it is
10− 6.
The simulation was run on an Intel Core2 quad-CPU Q6600 of
2.4 GHz. Three of the available four cores were used to run this model.
The simulation takes 15,000 iterations to converge, each iteration re-
quiring about 3 s to complete.

3. Results

The predicted steady-state results of the thermo-fluid performance


(hydrostatic pressure, flow patterns, thermal performance, vapour vo-
lume fraction and the mass transfer) inside the flashing chamber are
presented in Figs. 5 through 18.

3.1. Hydrostatic pressure distribution


Fig. 12. Temperature distribution along vertical traverses at x = 0.31 m, 0.62 m, 0.93 m.

Figs. 5, 6 and 7 show the hydrostatic pressure distribution


throughout the flashing chamber. A discontinuity in pressure gradient
is evident between the brine and the vapour layers, and the lower edge
of this discontinuity is taken to be the free surface. There is an essen-
tially linear increase in pressure with depth below the free surface.
Along a horizontal traverse at mid-orifice level (y = 0.03 m, Fig. 7) the
calculated pressure drop is 2179 Pa, corresponding to the pressure
difference required to drive the flow from the inlet to the outlet.

3.2. Particle paths and velocity distribution

Fig. 8 shows particle paths (streamlines), coloured by the speed of


the fluid mixture, throughout the flashing chamber. A detailed view of
the flow pattern in the inlet region is also provided. As commonly ob-
served in flashing chamber experiments [15–23], a distinct brine re-
circulation flow zone appears in the solution.
The vertical traverses (Fig. 9) indicate a sharp discontinuity in ve-
locity between the brine and vapour along the interface, with minima
occurring at the free surface. Along the mid-orifice level (Fig. 10) the
inlet flow decelerates due to jet spreading and subsequently accelerates
to a maximum of 1.88 m/s at the brine exit.

3.3. Thermal performance


Fig. 13. Temperature distribution along horizontal traverses at y = 0.03 m, 0.17 m, free
surface line, 0.3 m.
The rate of flashing is related to the thermal performance, or brine
temperature field. The flashing process across the flashing chamber

266
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

Fig. 14. Vapour volume fraction ( αv) fields with two different ranges of scale; (a) upper field range [0–1], and (b) lower field range [0–9 × 10− 5].

Fig. 15. Vapour volume fraction distribution along vertical traverses at x = 0.31 m, Fig. 16. Vapour volume fraction distribution along horizontal traverses at y = 0.03 m,
0.62 m, 0.93 m. 0.17 m, free surface line, 0.3 m.

267
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

vapour saturation pressure with depth, which affects the saturation


temperature. Thus mass transfer and phase change decrease with depth
due to the increase in pressure.
The hydrostatic pressure has the same effect as the brine level on the
flashing performance. The increase of the brine level leads to higher
thermodynamic non-equilibrium losses in the flashing chamber. Also,
the increase of the brine level leads to an increase in the pressure with
depth.
Generally, hydrostatic pressure [44] affects both the recirculation
zone shape and locations (it increases linearly with the brine level), and
bubble growth and rise locations. However, in our simulation case, the
hydrostatic pressure is not enough to block the bubbles formation and
the reduction in brine temperature.

4.2. Particle paths and the velocity distribution

Using the VOF in our simulations does not allow one to study the
velocity distribution separately for each phase in the mixture case. Thus
the velocity distributions in Figs. 8 to 10 are for the mixture only.
However, the VOF model makes it possible to deal with the vapour
layer as well with the brine layers.
Fig. 17. Vapour volume fraction distribution along horizontal traverses y = 0.03 m.
Vertically, the velocity magnitude at the free surface is the
minimum. This happens in order to keep conservation of mass concept
over the free surface valid. For x = 0.93 (Fig. 9) the value of the
causes a reduction of brine temperature, as can be seen in the overall mixture speed is higher comparing with x = and 0.31, 0.62 m. This is
temperature field (Fig. 11), and also in the vertical and horizontal due to the higher phase changes rate and thus more vapour phase in-
temperature traverses presented in Figs. 12 and 13. cluded in the mixture as it can be noticed in Fig. 14 at the lower field.
The decrease in temperature takes place in both the horizontal (x-) Horizontally, the maximum speed is found to be 1.88 m/s at the end of
and vertical (y-) directions, but at different rates. The differences in the flashing chamber. This happens to keep the conservation of energy
range and scale between Figs. 12 and 13 should be noted. The tem- concept valid, since pressure in this point has reached the minimum
perature gradient in vertical (y-) directions is large compared to the (x-) value.
horizontal. The calculated flashing down (Tin − Tout) is 4.55 °C. The recirculation zone promotes bubble transport from the bubble
nucleation region to the free surface. Other information about the re-
3.4. Vapour volume fraction and mass transfer rate lationship between fluid flow and its thermal behaviour can be reached
by using such simulations, i.e. rate at which turbulence kinetic energy
Fig. 14 presents the distribution of vapour volume fraction ( αv) converts into thermal internal energy. This can provide information
throughout the flashing chamber, and for clarity two field images are about vertical heat flow, which is required in order for evaporating li-
provided. The upper image (a) indicates the full range of vapour vo- quid close to the free surface.
lume fraction scale from 0 to 1, and shows that above the free surface
there is a mixture of vapour and liquid phases. The lower set of images
4.3. Thermal performance
(b) focuses on a restricted range, between 0 and about 0.0001, to better
illustrate the vapour phase distribution throughout the evaporation
Generally, thermodynamics of the brine in the evaporation zone
zone. The peak values of vapour volume fraction occur at the free
inside the flashing chamber is difficult to predict [70]. However, our
surface and in the bubble nucleation region. In the bubble nucleation
computational model for the evaporation zone can be used to obtain
region three different areas are numbered; the maximum bubble for-
more information about the thermodynamics of the brine inside the
mation takes place in area (1), in the inlet jet just below the top edge of
flashing chamber. We found that when the horizontal rate of decrease
the inlet orifice.
of temperature is greater than the vertical rate of decrease of tem-
Along vertical traverses above the free surface (Fig. 15), the vapour dT dT
perature ( dx > dy ), the non-equilibrium temperature difference (NETD)
volume fraction increase monotonically. Horizontally, (Fig. 16) the
dT dT
values of vapour volume fraction decrease. The average vapour volume will be positive, while for dx < dy , NETD will be negative. This will be
fraction along the free surface is 6.84 × 10− 3. Fig. 17 shows the dis- discussed in future work.
tribution of vapour volume fraction along y = 0.03 m, which passes The reduction in brine temperature in the evaporation zone occurs
through the bubble nucleation region. This happens because of the due to the change phase. This reduction in temperature affects the mass
reduction in brine temperature, so the phase change rate is decreased. transfer rate and vapour volume fraction distribution (see Figs. 14 to
Fig. 18 shows the mass transfer rate in the bubble nucleation region 18). Also the mixing process affects the rate of temperature decrease.
and at the free surface layer inside the flashing chamber. The positive In our simulation, the temperature gradient in the vertical (y-) di-
value means that the phase changes from liquid to vapour phase (va- rection is large compared to the (x-) horizontal. The turbulent shear
pourisation process is occurred) while the negative value means the mixing between the upper and lower brine layers influences the local
phase changes from vapour to liquid phase (condensation process is temperature gradient. The upper layer is colder than the lower, and
occurred). since the horizontal temperature gradient is small compared to the
vertical gradient, the preferred direction of heat transfer is vertical
4. Discussion rather than horizontal.
In our simulation we found that the driving force of the phase
4.1. Hydrostatic pressure distribution change — the temperature difference between the brine temperature
and the saturation temperature — along the evaporation zone has re-
Hydrostatic pressure resists the flashing process [44] by increasing mained positive. In other words, while the phase change reduces the

268
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

Fig. 18. Mass transfer rate field with three different ranges of scale; (a) [− 190, + 34] kg/m3s, (b) [− 26, + 8] kg/m3s, and (c) [− 0.0016, + 2.5] kg/m3s.

Table 1 occurrence of finite flashing? Does this mean that the flashing process
Comparison of predicted results and measured values. can be classified based on finite and infinite process? This is worthy of
investigation, and will be considered in future work.
Measured [46] Predicted

Average outlet temperature [K] 379 378.6


Average vapour temperature [K] 375 377.7 4.4. Vapour volume fraction and mass transfer rate
Total flashed vapour [tonne/h] 12.5 13.6

The vapour volume fraction distribution can be used as an indicator


for tracking the vapour phase and bubble formation.
brine temperature, its values are still higher than the saturation tem-
The maximum value of vapour volume fraction in the evaporation
perature corresponding to the flashing chamber pressure, thus main-
zone is 74 × 10− 6 (Fig. 17) and it is located in the bubble nucleation
taining the phase change along the evaporation zone length. This raises
region. The small value obtained is due to hydrostatic pressure (brine
the question of what if, at some location along the evaporation zone,
level) effects. On the other hand, the average value of vapour volume
the phase change reduces the brine temperature to the point where it
fraction at the free surface is 6.84 × 10− 3, which is higher than found
reaches the saturation temperature or lower? Does this lead to no more
in the bubble nucleation region. Thus phase change and mass transfer
change of phase taking place inside the evaporation zone, and the
occurs predominantly at the free surface (Fig. 18), similar to results

269
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

Fig. 19. Flashing regions as predicted by computational model. Compare with Fig. 2.

observed in the experiments of Gilbert [15]. vapour mass flow rate (as noted above in §4.4). In comparing our re-
The total vapourisation mass flow rate can be calculated by com- sults with the available data [46], we find a good agreement with the
puting the net produced vapour flux through the open boundary. From values for the real plant (Table 1).
the field results along the open boundary, we obtain a mean product of The predicted flashing mechanism in the simulation agrees with the
y-component velocity times average fluid density of 0.433 kg/m2 s. The mechanism described in §1.2 (see Fig. 2). As seen in Fig. 19 vapour
flashing chamber length is 1.25 m, so that for the corresponding width bubbles are formed at the entrance to the flashing chamber, and bubble
of 7 m [47], the estimated vapour mass flow rate is 3.79 kg/s production diminishes along its length. At the free surface of the liquid
(13.6 tonne/h). there is also a phase change and mass transfer. Thus the two main
mechanisms of phase change produced by the flashing process are
captured in the simulation.
4.5. Validation
The predicted level of brine inside the flashing chamber is 0.2 m
above the inlet. This agrees with design recommendations by El-
The main challenges for doing such predictions lie in obtaining
Dessouky et al. [6], who recommended that the brine pool be higher
sufficient data to build and to validate the computational model. In
than the gate height by 0.2 m.
particular, it is difficult to find information pertaining to the interior of
The calculated non-equilibrium temperature difference (NETD) va-
the flashing chamber, i.e. bubble nucleation rate, bubble formation,
lues range from 4.8 to 5.5 °C, which is found to fall within the range of
recirculation zone size and length, brine level, and also the shapes of
design recommendations [73]: the NETD for a flashing chamber
the orifices and their number. Also, for the few data that are available,
without a baffle lies between 4 and 5 °C.
there is little or no information about measurement locations, methods
or uncertainties.
The VOF [51] multiphase model is well established, and has been 4.6. Future work
validated for predicting liquid-gas interfaces, including the free surface
shape and levels, in a variety of other applications [51,52,59]. Tami- From the present work it is clear that computational field modelling
larasan [62] has validated the mechanistic Lee Wen Ho vapourisation- can be invaluable for a better understanding of the combined heat
condensation model for vapour volume fraction prediction, including transfer, mass transfer and the fluid dynamics during the flashing flow
its application within the Fluent VOF implementation. We have in- evaporation processes. We note that our method is capable of predicting
dependently validated the implementation of the Zwart et al. vapour- the fully three-dimensional, time-dependent behaviour of the eva-
isation-condensation model within the Fluent VOF code using data from poration zone, which exhibits oscillating, unsteady flow phenomena
detailed and extensive experiments for isothermal flashing water flow characteristic of open channel flows. While feasible, for such calcula-
in a converging-diverging nozzle by Abuaf et al. [71]. Under non- tions the computational time and cost is very high: moreover, assim-
equilibrium (flashing) conditions, the measured average vapour volume ilating and presenting the copious results of a transient analysis is time-
fraction at the outlet of the nozzle was 0.544, for which case our pre- consuming and challenging. Despite this potential, as already noted in
dicted value is 0.543 [72]. §4.5 the key obstacle nonetheless remains the lack of measurements
Our simulation of the evaporation zone of the flashing chamber is and data for validation.
validated by checking values of average temperature at the stage exit, In the interim, the current model, run in a steady, two-dimensional
average vapour temperature above the liquid free surface and the mode, will be used for studying and comparing the effects of variations

270
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

in the brine temperature, velocity and flashing chamber geometry on Flashing Salt Solution in Open Channel Flow Between Two Submerged Orifices at
Low Pressure (Ph.D. Thesis), University of Connecticut, Connecticut, 1968.
the thermo-fluid behaviour inside the evaporation zone. [16] O. Miyatake, K. Murakami, Y. Kawata, T. Fujii, Fundamental experiments of flash
The flashing process may be classified as a finite or infinite evaporation, Heat Transf. Jpn. Res. 26 (1972) 189–198.
(flashing) process. Depending on which of the two flashing process [17] T. Fujii, O. Miyatake, T. Tanaka, T. Nakaoka, H. Matsunaga, N. Sakaguchi,
Fundamental experiments on flashing phenomena in a multi-stage flash evaporator,
variations occurs — finite or infinite —may determine the thermo-fluid Heat Transf. Jpn. Res. 1 (1975) 207–212.
behaviour inside the evaporation zone inside the flashing chamber, and [18] O. Miyatake, T. Fujii, T. Tanaka, T. Nakaoka, Flash evaporation phenomena of pool
will also be investigated in future work. water, Heat Transf. Jpn. Res. 1 (1975) 393–398.
[19] O. Miyatake, T. Fujii, T. Hashimoto, An experimental study of multi-stage flash
evaporation phenomena, Heat Transf. Jpn. Res. 1 (1975) 454–459.
5. Conclusions [20] M.A. Darwish, S.G. Serageldin, H.T. El-Dessouky, Flashing in desalination units,
Desalination 19 (1976) 93–101.

• A VOF two-phase computational model of the evaporation zone [21] O. Miyatake, T. Nakaoka, T. Fujii, The effect of nucleation on flash evaporation,
Heat Transf. Jpn. Res. 29 (1976) 228–231.
inside a flashing chamber has been developed. The model includes [22] T. Sugeta, S. Toyama, An experimental study of flash evaporation phenomena at
heat transfer and mass transfer, as well as turbulent flow. high liquid flow rate, Heat Transf. Jpn. Res. 4 (1978) 148–153.

• Two different phase-change mechanisms are allowed for, based on [23] N. Lior, R. Greif, A system for the experimental study of flash evaporation,
Desalination 31 (1979) 87–99.
the saturation temperature and the vapour pressure, respectively. [24] N. Lior, E. Nishiyama, The effect of gas bubbles on flash evaporation, Desalination
These enable the model to compute the phase change regions, also 45 (1983) 231–240.
the shape of the free surface. [25] N. Lior, Formulas for calculating the approach to equilibrium in open channel flash


evaporators for saline water, Desalination 60 (1986) 223–249.
The model has been applied to a typical horizontal-flow flashing [26] S. Gopalakrishna, V.M. Purushothaman, N. Lior, An experimental study of flash
chamber without a flow baffle. In a preliminary, steady-state solu- evaporation from liquid pools, Desalination 65 (1987) 139–151.
tion the flow field features, including the recirculation zone and the [27] S. Gopalakrishna, N. Lior, Analysis of bubble translation during transient flash
evaporation, Int. J. Heat Mass Transf. 35 (1992) 1753–1761.
free surface level and shape are predicted. The simulation results [28] O. Miyatake, Comparative study of flashing evaporation rate, Desalination 96
give average outlet temperature and average vapour temperature (1994) 163–171.
values that agree closely with those of a real MSF plant. [29] N. Lior, E. Nishiyama, Some experiments on flash evaporation enhancement by

• Computational field modelling is valuable for a better under-


[30]
electrolytically generated bubbles, Desalination 100 (1995) 71–76.
O. Miyatake, I. Tanaka, N. Lior, A simple universal equation for bubble growth in
standing of the combined heat transfer, mass transfer and the fluid pure liquids and binary solutions with a nonvolatile solute, Int. J. Heat Mass Transf.
dynamics during the flashing flow evaporation processes. 40 (1997) 1577–1584.

• Results can be used to estimate MSF design factors such as the non- [31] J.-I. Kim, N. Lior, Some critical transitions in pool flash evaporation, Int. J. Heat
Mass Transf. 40 (1997) 2363–2372.
equilibrium temperature difference, NETD, flashing down and [32] H.E.S. Fath, The non-equilibrium factor and the flashing evaporation rate inside the
flashing efficiency. flash chamber of a multi-stage flash desalination plant, Desalination 114 (1997)

• The model can be applied for further studies such as investigations


[33]
277–287.
W.X. Jin, S.C. Low, S.C.M. Yu, Some experimental observations on the single and
of the effects of the brine temperature, velocity and flashing multi-phase flow patterns in a model flash evaporation chamber, Int. Commun.
chamber geometry on the thermo-fluid behaviour for both finite and Heat Mass Transf. 26 (1999) 839–848.
infinite flashing process classifications. [34] O. Miyatake, Y. Koito, K. Tagawa, Y. Maruta, Transient characteristics and per-
formance of a novel desalination system based on heat storage and spray flashing,
Desalination 137 (2001) 157–166.
Acknowledgments [35] W.X. Jin, S.C. Low, Investigation of single-phase flow patterns in a model flash
evaporation chamber using PIV measurement and numerical simulation,
Desalination 150 (2002) 51–63.
This work was funded through a postgraduate Research Fellowship [36] X. Shichang, L. Wencai, W. Shichang, Experimental studies on orifice behavior and
from the College of Engineering and Informatics, National University of two-phase flow in flash chamber, Desalination 150 (2002) 93–98.
Ireland, Galway, Ireland. [37] N. Lior, R. Chung, O. Miyatake, Correlations (updated) for predicting the flow
through MSF plant interstage orifices, Desalination 151 (2003) 209–216.
[38] D. Saury, S. Harmand, M. Siroux, Flash evaporation from a water pool: influence of
References the liquid height and of the depressurization rate, Int. J. Therm. Sci. 44 (2005)
953–965.
[39] A. Cipollina, Experimental Study and Dynamic Modelling of Multi Stage Flash
[1] A.H. Khan, Desalination Processes and Multi-stage Flash Distillation Practice,
Desalination Units (Ph.D. Thesis), Università di Palermo, Palermo, 2005.
Elsevier, Amsterdam, 1986.
[40] A. Cipollina, G. Micale, L. Rizzuti, Investigation of flashing phenomena in MSF
[2] L.K. Wang, J.P. Chen, Y. Hung, N.K. Shammas, Membrane and Desalination
chambers, Desalination 216 (2007) 183–195.
Technologies, Springer, London, 2011.
[41] K.W. Seul, S.Y. Lee, Numerical predictions of evaporative behaviour of horizontal
[3] Fichtner GmbH, MENA Regional Water Outlook Part II: Desalination Using
stream inside a multi-stage-flash distillator, Desalination 79 (1990) 13–35.
Renewable Energy, Report, Germany, (2011).
[42] C.T. Crowe, M.P. Sharma, D.E. Stock, The particle-source-in cell (PSI-CELL) method
[4] International Desalination Association, IDA Desalination Yearbook 2013–2014,
for gas droplet flow, J. Fluids Eng. 99 (1977) 325–332.
Media Analytics Ltd., Oxford, United Kingdom, 2014.
[43] O. Miyatake, T. Hashimoto, N. Lior, The liquid flow in multi-stage flash evapora-
[5] F.A. Al-Marzooqi, A.A. Al-Ghaferi, I. Saadat, N. Hilal, Application of capacitive
tors, Int. J. Heat Mass Transf. 35 (1992) 3245–3257.
deionisation in water desalination: a review, Desalination 342 (2014) 3–15.
[44] K.W. Seul, S.Y. Lee, Effect of liquid level on flow behaviour inside a multi-stage
[6] H.T. El-Dessouky, H.M. Ettouney, Fundamentals of Salt Water Desalination, Elsevier
flash evaporator - a numerical prediction, Desalination 85 (1992) 161–177.
Science B.V., Amsterdam, the Netherlands, 2002.
[45] O. Miyatake, T. Hashimoto, N. Lior, The relationship between flow pattern and
[7] D.W. Abueidda, M. Gaddalla, Thermodynamics analysis of integrated solar-based
thermal non-equilibrium in the multi-stage flash evaporation process, Desalination
multi-stage flash desalination system, ASME International Mechanical Engineering
91 (1993) 51–64.
Congress & Exposition, IMECE2012-87729, Nov 2012.
[46] K.M. Mansour, H.E.S. Fath, Numerical simulation of flashing process in MSF flash
[8] M. Al-Weshahi, Exergy Efficiency Enhancement of MSF Desalination by Heat
chamber, Desalin. Water Treat. 51 (2013) 2231–2243.
Recovery From Hot Distillate Water Stages, Newcastle University, UK, 2011.
[47] K.M. Mansour, H.E.S. Fath, O. El-Samni, Computational fluid dynamics study of
[9] N. Lior, Some basic observations on heat transfer and evaporation in the horizontal
MSF flash chambers sub-components; I – vapor flow through demister, The
flash evaporator, Desalination 33 (1980) 269–286.
Fifteenth International Water Technology Conference, IWTC15. Alexandria, Egypt,
[10] R. Rautenbach, S. Schäfer, S. Schleiden, Improved equations for the calculation of
2011.
non-equilibrium temperature loss in MSF, Desalination 108 (1996) 325–333.
[48] ANSYS Fluent Theory Guide, Release 14.5, (2016).
[11] B.M. Fabuss, A. Korosi, Boiling point elevations of sea water and its concentrates, J.
[49] J.H. Ferziger, M. Peric, Computational Methods for Fluid Dynamics, Springer, New
Chem. Eng. Data 11 (October 1966) 606–609.
York, 2002.
[12] AMF, Cuno Engineering Corporation Maximum Division, Research and
[50] W.F. Noh, P. Woodward, SLIC (Simple Line Interface Calculation), in: A.I. van de
Development of Multieffect, Multistage Flash Distillation Process, OSW R & D
Vooren, P.J. Zandbergen (Eds.), Proceedings, Fifth International Conference on
Annual Report 15 (October 1966).
Fluid Dynamics, Lecture Notes in Physics, 59 Springer, Berlin, 1976, pp. 330–340.
[13] Westgrath W. Ltd., Chamber Geometry in Multi-stage Flash Evaporator-separator
[51] B.D. Nichols, C.W. Hirt, Numerical simulation of boiling water reactor vent-clearing
Studies, OSW R & D Progress Report No. 271, (July 1967).
hydrodynamics, Nucl. Sci. Eng. 73 (1980) 196–209.
[14] W.R. Williamson, F.W. Gilbert, Investigation of Multi-stage Flash Phenomena in a
[52] C.W. Hirt, B.D. Nichols, Volume of fluid (VOF) method for the dynamics of free
Three-stage Test System, OSW R & D Progress Report No. 525, (Feb 1968).
boundaries, J. Comp. Phys. 39 (1981) 201–225.
[15] F.W. Gilbert, An Experimental Investigation of Steady State Characteristics of

271
T.H. Nigim, J.A. Eaton Desalination 420 (2017) 258–272

[53] J.M. Hyman, Numerical methods for tracking interfaces, North-Holland Physics [64] W.H. Lee, A Pressure Iteration Scheme for Two-phase Modeling, Technical Report
Publishing Division, Physica 12D (1984) 396–407. LA-UR 79-975, Los Alamos Scientific Laboratory, Los Alamos, New Mexico, 1979.
[54] O. Ubbink, Numerical Prediction of Two Fluid Systems With Sharp Interfaces (Ph.D. [65] P.J. Zwart, A.G. Gerber, T.A. Belamri, Two-phase flow model for predicting cavi-
Thesis), University of London, London, 1997. tation dynamics, Fifth International Conference on Multiphase Flow, Yokohama,
[55] W.J. Rider, D.B. Kothe, Reconstructing volume tracking, J. Comp. Phys. 141 (1998) Japan, 2004.
112–152. [66] C.E. Brennen, Fundamentals of Multiphase Flow, Cambridge University Press, 2005.
[56] R. Scardovelli, S. Zaleski, Direct numerical simulation of free surface and interfacial [67] N.B. Vargaftik, B.N. Volkov, L.D. Voljak, International tables of the surface tension
flow, Ann. Rev. Fluid Mech. 31 (1999) 567–603. of water, J. Phys. Chem. 12 (1983) 817–820.
[57] S. Kvicinsky, F. Longatte, J.L. Kueny, F. Avellan, Free Surface Flows: Experimental [68] M.H. Sharqawy, J.H. Lienhard V, S.M. Zubair, Thermophysical properties of sea-
Validation of Volume of Fluid (VOF) Method in the Plane Wall Case, Proceedings of water: a review of existing correlations and data, Desalin. Water Treat. 16 (2010)
the 3rd ASME, JSME Joint Fluids Engineering Conference, San Francisco, 354–380.
California, (1999). [69] H. Versteeg, W. Malalasekera, An Introduction to Computational Fluid Dynamics:
[58] R. Saurel, R.A. Abrall, Multiphase Godunov method for compressible multifluid and The Finite Volume Method, 2nd ed., Prentice Hall, 978-0-13-127498-3, 2007.
multiphase flows, J. Comp. Phys. 150 (1999) 425–467. [70] R. Rautenbach, S. Schafer, Inter-stage Brine Flow in MSF Chambers, Desalination
[59] D.M. Hargreaves, H.P. Morvan, N.G. Wright, Validation of the volume of fluid and Water Resources: History, Development and Management of Water Resources
method for free surface calculation: the broad-crested weir, Eng. Appl. Comput. vol. I, (DESWARE). United Nations Educational, Scientific and Cultural
Fluid Mechanics 1 (2007) 136–146. Organization (UNESCO), & Encyclopaedia of Life Support Systems (ELSS), 978-1-
[60] V. Hernandez-Perez, Gas-liquid Two-phase Flow in Inclined Pipes (PhD thesis), 84826-419-9, 2010.
University of Nottingham, 2008. [71] N. Abuaf, B.J.C. Wu, G.A. Zimmer, P. Saha, A Study of Non-Equilibrium Flashing of
[61] A.P. Laleh, W.Y. Svrcek, W.D. Monnery, Design and CFD studies of multiphase se- Water in a Converging-Diverging Nozzle: Volume 1 – Experimental, United States
parators – a review, Can. J. Chem. Eng. 90 (2012). Nuclear Regulatory Commission Office of Nuclear Regulatory Research, 1981.
[62] A. Tamilarasan, Numerical Modelling of Heat Transfer and Evaporation [72] T.H. Nigim, Computational Modelling of Thermofluid Flashing in MSF Desalination
Characteristics of Cryogenic Liquid Propellant (Master of Science in Mechanical (Ph.D. Dissertation), National University of Ireland, Galway, 2017 (submitted for
Engineering Thesis), University of Washington, 2015. publication).
[63] K.M. Wrobel, Modelling and Numerical Simulation of Flows With Evaporation and [73] Desalting Systems and Services, Inc., Design and Furnishing Flash Enhancers for
Condensation (Ph.D. Thesis), University of Twente, Enschede, 90-365-2387-7, MSF Test Module, OSW Res. Devel. Progress Report No. 565, (1971).
2006.

272

You might also like