You are on page 1of 9

Desalination 259 (2010) 29–37

Contents lists available at ScienceDirect

Desalination
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / d e s a l

Simulation and optimisation of direct contact membrane distillation for


energy efficiency
V.A. Bui a,1, L.T.T. Vu b, M.H. Nguyen c,⁎
a
Centre for Plant and Food Science, University of Western Sydney, NSW, Australia
b
School of Engineering and Energy, Murdoch University, WA, Australia
c
School of Natural Sciences, University of Western Sydney and School of Environmental and Life Sciences, University of Newcastle, NSW, Australia

a r t i c l e i n f o a b s t r a c t

Article history: This paper describes the formulation of a computational framework for simulating and optimising DCMD to
Received 11 February 2010 minimise the consumed energy. A simulation procedure for DCMD was established on the basis of equating
Received in revised form 15 April 2010 heat and mass fluxes through different domains in the process. Steady-state simulations for a wide range of
Accepted 16 April 2010
operating conditions were carried out. It was revealed that the highest achievable energy efficiency of DCMD
Available online 15 May 2010
within the tested range was about 49.9%.
A double loop optimisation problem was formulated in MATLAB to solve the highly nonlinear equations with
Keywords:
Direct contact membrane distillation
unknown outlet and membrane surface conditions to implement the simulation procedure. An additional
DCMD outer loop was also implemented to accommodate the dynamic condition of a real lab-scale DCMD system
Energy efficiency concentrating 1.5 kg glucose solution from 30 to 60% w/w. A pseudo-real-time dynamic optimisation was
Simulation performed to minimise the energy expenses for the DCMD process. This energy accounted for the heat
Optimisation exchanged between the feed and permeate streams within the membrane module and the power for their
pumping, while maintaining a minimum mass flux of 0.5 kg m− 2 h− 1. The optimal operating conditions
found in this study could save the total energy consumption by 26.3%.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction heat transfers across the membrane in DCMD. Consequently, as in any


transport process, multiple polarisations exist concurrently in DCMD
Direct contact membrane distillation (DCMD) is a selective as illustrated in Fig. 1. All symbols and abbreviations used in this
membrane separation process driven by a vapour pressure gradient figure and throughout this paper are in the nomenclature list.
across a hydrophobic microporous membrane, which stays as a barrier DCMD is applicable in a wide range of areas but has been mostly
between a hot feed and a cold permeate. The hydrophobic studied in dewatering applications such as desalination, waste water
characteristic of the membrane prevents any liquid from entering treatment and concentration of liquid solutions [2–4]. DCMD has been
the pores unless the membrane is under a hydrodynamic pressure widely recognised as cost-efficient for desalination and waste water
higher than its liquid penetration point (LPP), creating gas–liquid treatment operating at temperatures from 60 to 90 °C when waste
interfaces on both sides of the membrane. Air that has filled the heat is employed to power the process [5,6]. Recently, DCMD has been
membrane pores stays immobilised throughout the DCMD operation. of interest in liquid food concentration application for its ability to
Once a water vapour pressure (WVP) gradient exists across the operate at temperatures as low as of 25 to 40 °C, yet achieving high
membrane, water molecules start to evaporate from the gas–liquid solid content, up to 65% w/w [7–11]. From this point of view and as a
interface on the feed side, diffuse through the immobilised air in the chemical-free process, DCMD is apparently advantageous over
pores, condense on the permeate side, and eventually are swept away osmotic distillation, a process that has attracted many researchers
with the permeate stream [1]. over the last two decades. Application of DCMD as such is envisaged as
The temperature difference between the two streams is the main a subsequent and final concentration step following reverse osmosis
factor that creates the driving force for DCMD mass transfer. Thus, or freeze concentration for thermo-sensitive liquid products.
there co-exist WVP and temperature gradients, and thereby mass and However, low temperature DCMD has encountered a serious
challenge of energy inefficiency as shown in a number of studies [12–14],
which poses a serious drawback to its commercialisation in the
industry. There is a need to optimise the process in order to minimise
⁎ Corresponding author. Fax: +61 2 4570 1369.
E-mail address: m.nguyen@uws.edu.au (M.H. Nguyen).
the required energy, while maintaining an adequate flux throughout
1
Current address: Faculty of Food Science and Technology, Nong Lam University, the operation. Such a goal can only be achieved by computer-aided
Vietnam. simulation and optimisation.

0011-9164/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2010.04.041
30 V.A. Bui et al. / Desalination 259 (2010) 29–37

Fig. 1. Principle of DCMD and the associated polarisations.

This paper presents the formulation of a computational framework loss kins ⋅Alm 1
Q = ðTf −Tamb Þ = ðT −Tamb Þ: ð4Þ
for solving the highly nonlinear equation system to simulate the δins 1:426 f
DCMD transport processes within a hollow fibre module (HFM). A
batch process was then optimised for minimal energy expense to
tackle the problem of high energy demand in low temperature DCMD The constant used in Eq. (4) was calculated based on given values
applications. of the physical properties of the insulation pipe. They included
thermal conductivity k ins = 0.26 W m − 1 K − 1 , inner diameter
2. Problem description and formulation 12.7 mm, thickness δins = 8 mm, and length L = 350 mm.
Furthermore, the heat flux across the membrane (Qm) comprises
Fig. 2 shows the heat fluxes dependent on inlet and outlet the latent heat carried by the water vapour diffusing across the
temperatures in a counter-current flow DCMD. The energy balance membrane (Qdiff) and the heat of conduction (Qcond) as shown in
requires: Eq. (5). This heat flux must equal to convective heat flux through the
feed boundary layer, (Qf), and to that through the permeate boundary
P F loss
Q = Qm = Q −Q : ð1Þ layer, (Qp), as respectively defined by Eqs. (6) and (7). Thus, another
heat balance was formed as given in Eq. (8).
In Eq. (1), heat released by the feed (QF), heat gained by the
 
permeate (QP) and heat loss through the module insulation (Qloss) Q m = Q cond + Q diff = Ai ⋅hm ⋅ Tfm −Tpm + Ai ⋅J⋅ΔHv ð5Þ
were respectively defined by Eqs. (2)–(4).

F
Q = ṁf ⋅Cpf ⋅ðTfi −Tfo Þ ð2Þ Q f = Ao ⋅hf ⋅ðTf −Tfm Þ ð6Þ

   
P
Q = ṁp ⋅Cpp ⋅ Tpo −Tpi ð3Þ Q p = Ai ⋅hp ⋅ Tpm −Tp ð7Þ

Fig. 2. Corresponding heat fluxes due to temperature changes from inlets to outlets in a counter-current flow DCMD (a); and the hypothesised heat balance at correct outlet
permeate temperature (b). (QF, heat released by the feed; QP, heat gained by the permeate; Qloss, heat lost through the HFM's insulation; Qm, heat transferred across the membrane;
Qf, convective heat flux through feed boundary layer; Qp, convective heat flux through permeate boundary layer; Tpi, Tpo, permeate inlet and outlet temperatures; Tfi, Tfo, feed inlet
and outlet temperatures).
V.A. Bui et al. / Desalination 259 (2010) 29–37 31

Q f = Q p = Q m: ð8Þ Pfm by an order of its corresponding water activity as compared to the


water vapour pressure of pure water at the same temperature.
Meanwhile, the membrane vapour permeability (Km) is dependent on
J in Eq. (5) is the mass flux based on the internal surface area. To membrane physical properties such as tortuosity, porosity, and
calculate Qm, J must be known beforehand. Therefore an iterative diffusivity of water vapour through stagnant air. Values of Km used
optimisation problem was developed for this purpose. An interme- in this work were determined based on the molecular diffusion
diate mass flux JT was derived in Eq. (9) from an analytical solution of model, using data obtained from experiments [18]. This mechanism of
the equation system, from Eqs. (5)–(8). vapour transport has been successfully applied in a number of studies
in DCMD on a flat sheet membrane [27–30] and was modified for
h   i   cylindrical geometry of the hollow fibres in this work. Specifically,
ðTf −Tfm Þ⋅ hm + 1 + hm = hp ⋅hf ⋅ðro = ri Þ − Tf −Tp ⋅hm thickness of the flat sheet membrane in the model was replaced by
JT = : ð9Þ [ri / log(ro / ri)] when calculating Km when based on the internal
ΔHv
surface area of the hollow fibres.
Values of JC calculated from Eq. (13) is substituted in Eq. (5) to
calculate Qm. The formulated problem is converged when the
To calculate the heat transfer coefficient (hp) in the lumen side for difference between the heat fluxes Qm and Qp or Qm and Qf are
water, Sieder–Tate or Hausen heat transfer models were applied for minimal. Similarly at convergence the deviation between the mass
laminar flow within a cylindrical pipe (Re b 2100 in this study). These fluxes JT and JC are less than a selected tolerance, e.g. 10− 6. These mass
models can be found in heat transfer textbooks [15,16], with the fluxes are used throughout the calculation procedure only in the steps
viscosity factor included [17]. To calculate the heat transfer (hf) and specifically described by Eqs. (9) and (13), respectively.
mass transfer (Kf) coefficients in the shell side for the feed, models in Experiments were performed as described in the following section
the form of Eqs. (10) and (11) were applied with the constants to measure the outlet temperatures and the mass flux of water.
according to Bui [18]. These constants are listed in Table 1 for the five
modules described in Section 3. 3. Experimental setup
 φ  
α β d μ f 0:14
Nuf = B⋅Ref ⋅Prf ⋅ hf ⋅ ð10Þ Operation of the DCMD process was configured within a hollow
L μ fm
fibre module (HFM) in a counter-current flow arrangement, shown in
 φ   Fig. 3, to concentrate model glucose solutions from 30 to 60% w/w. In
α β dhf μ f 0:14
Shf = B⋅Ref ⋅Scf ⋅ ⋅ μ : ð11Þ this arrangement, the feed, hot glucose solutions at temperatures in
L fm the range of 25–40 °C, was allowed to flow in the shell side at
velocities of 0.4–0.8 m s− 1; and the permeate, cold water at
The thermal conductivity of the porous membrane (hm) could be temperatures lower than those of the feed, in the lumen side at the
estimated by using a theoretical model or a regression analysis of velocity of 1.2 m s− 1. These velocities were controlled by adjusting
experimental data [18]. valves 11 and 12. The upper limits were put on the feed and permeate
The mass flux in Eq. (9) was only dependent on the heat transfer in flow rates to avoid liquid penetration through membranes. Five HFMs
a balanced DCMD system, rather than being governed by the made of two different materials were tested. The specifications are
molecular diffusion law. Therefore, it was used in the calculation of shown in Table 2, with PV for polyvinylidene fluoride (PVDF) and HL
the feed concentration at the membrane surface, as in Eq. (12). This for ethylene-chlorotrifluoroethylene copolymer (Halar). The hollow
equation was based on the ‘thin film’ theory, which has been validated fibres were supplied by Siemens Water Technology (South Windsor,
in a number of numerical studies on DCMD and reverse osmosis [19–21]. NSW, Australia).
This equation has been extensively used in the mass transfer analysis of During the experiments, the amount of water removed from the
membrane distillation [22,23]; and osmotic distillation [24–26]. feed into the permeate over every 15 min was precisely measured by
  the water level rising in pipette 13, and the mass flux was calculated.
JT ⋅ri = ro
Cfm = Cf ⋅ exp ð12Þ This amount of water was then discharged via valve 14. Temperature
ρf ⋅Kf

On the other hand, Eq. (13) was deployed to calculate the mass
flux JC, from the feed to the permeate.
 
JC = Km ⋅ Pfm −Ppm : ð13Þ

The water vapour pressures at the membrane surfaces (Pfm) and


(Ppm) can be obtained from the Antoine's equation, providing that the
corresponding temperatures at the membrane surfaces have been
calculated in each loop of the iterative optimisation problem. It should
be noted that glucose concentration at the membrane surface reduces

Table 1
Constants in the semi-empirical models given in Eqs. (10) and (11).

Module B̅ α β φ

PV80 7.03 0.63 0.38 0.96


PV65 6.42 0.75 0.39 1.09
PV37 4.39 1.06 0.55 1.16 Fig. 3. Schematic diagram of the lab-scale DCMD unit using HFMs. (1, HFM; 2, feed
HL50 2.83 0.64 0.45 0.95 bottle; 3, permeate bottle; 4 and 5, pumps; 6, hot water bath; 7, cool water bath; 8, flow
HL31 2.41 0.68 0.42 0.89 meters; 9, pressure gauges; 10, thermometers; 11 and 12, flow regulating valves; 13,
flux recording pipette; 14, permeate discharge valve).
32 V.A. Bui et al. / Desalination 259 (2010) 29–37

Table 2
Specifications of the hollow fibre modules used in DCMD experiments.

Code n ϕ di L LT Ai Al Ash dh
% mm mm mm cm2 mm2 mm2 mm

PV37 139 50.4 0.37 350 400 573.1 15.4 42.0 0.550
PV65 54 50.1 0.65 350 400 385.9 17.9 42.2 0.835
PV80 32 50.2 0.80 350 400 281.5 16.1 42.2 1.032
HL31 128 50.2 0.31 350 400 436.3 9.66 42.2 0.573
HL50 85 50.5 0.50 350 400 467.3 16.7 41.9 0.681

n, number of fibres; ϕ, packing density; di, internal diameter of fibres; L, effective mass
transfer length; LT, total length; Ai, internal mass transfer area; Al, Ash, cross-sectional
areas in the lumen side and in the shell side, respectively; dh, equivalent hydrodynamic
diameter of the shell side flow channel.

changes of the feed and the permeate streams from the inlets to the
outlets were accordingly recorded in each run for calculations of heat
exchange rates in the process.

4. Problem solving methods

4.1. Steady-state simulation

As the heat and mass transfer models are dependent on physical


properties of the feed and permeate streams, the outlet temperature
of either one of these streams (Tfo or Tpo) was initially selected in the
first iteration loop. Temperatures at the other stream's outlet and of
the membrane module were then calculated, and the physical
properties estimated. The amount of heat transferred between two
streams (QF, QP) and the heat flux through the membrane (Qm) could
be calculated. The optimisation problem was then set up with the
objective function to minimise the square of the deviation of the heat
fluxes, min[(QP − Qm)2]. The new value found for Tfo was fed to the
second iteration of the iterative optimisation problem. The same
procedure was repeated until the differences between calculated
variables from the current iteration and the previous one were less Fig. 4. Algorithm for steady-state simulation of DCMD in an HFM. (Tmf, Tmp, Cmf, initially
than the selected tolerance, e.g. 10− 6. The problem was programmed guessed or temporary feed and permeate temperatures and feed concentration at the
membrane surfaces, respectively; Tfm, Tpm, Cfm, the corresponding calculated tempera-
and solved using MATLAB and its optimisation Toolbox following the
tures and concentration).
algorithm shown in Fig. 4.

4.2. Dynamic simulation of circulated DCMD operation

A DCMD lab-scale unit as described in Fig. 3 using HFM can only


work in a batch mode with the feed stream recirculated within the
module for a certain amount of time before its concentration reaches
the desired value Cf(tf). The feed concentration increases from the
initial Cf(t0) due to the removal of water over the course of the
concentration operation was calculated by Eq. (14).

Cf ðt0 Þ⋅M0
Cf ðt Þ = t : ð14Þ
M0 −Ai ⋅ ∫ J ðt Þdt
0

To formulate this dynamic simulation, an outer loop was added


to the iterative steady-state simulation problem, as shown in Fig. 5.
It is obvious that this outer loop was performed once the steady-
state simulation problem had been converged. As a result, either JT
or JC could be used in the place of J(t) in Eq. (14) due to their same
values.
The dynamic problem was simulated to determine the total
processing time and energy required for concentrating an amount of
feed solution from the initial concentration Cf(t0), to the final Fig. 5. Algorithm for dynamic simulation of DCMD in an HFM. (Steady-state loop refers
concentration Cf(tf). The total energy was then defined as a sum of to the algorithm in Fig. 4).
V.A. Bui et al. / Desalination 259 (2010) 29–37 33

the energy supplied to the feed and the energy for circulation of the
two streams in the system, and was calculated in Eq. (15).
tf  
F
EDCMD = ∫ Q + Qpump dt ð15Þ
0

The energy consumed by the pump, (Qpump), was a function of the


fluid mass flow rates (ṁ), viscosity and the flow geometry. Eq. (16)
was used to calculate this power, assuming that the pressure lost due
to the piping could reach around 15% of the pressure drop through the
membrane Δpj.

j
Q pump = ∑ Q pump = ∑ ṁj ⋅ð1 + 0:15Þ⋅Δpj j = feed; permeate
j j

ð16Þ
Fig. 7. Comparison between simulated and experimental mass fluxes.

4.3. Dynamic optimisation of circulated DCMD operation


heat and mass fluxes in modules HL50 and PV65 are illustrated in
To optimise a DCMD operation, various objective functions can be Figs. 7 and 8, and comparisons for all modules are listed in Table 3.
selected such as to minimise the processing time, or to minimise the According to the ratios in Table 3, the simulated mass fluxes
total energy (EDCMD) defined in Eq. (15), or to maximise the total statistically differed from the experimental ones by only 0.50% [for
energy efficiency (EEDCMD) defined in Eq. (17). PV65] to 3.24% [for HL31], while those for heat fluxes varied between
0.03% [for PV37] and 2.39% [for HL50]. These negligible differences
t
Ai ⋅ ∫ J ðt Þ⋅ΔHv ðt Þdt and the coefficients of determination, R2, that were above 0.95
suggested that the new models [18] successfully described heat and
EEDCMD = 0
ð17Þ
EDCMD mass transfer in the DCMD process, reproducing both heat and mass
fluxes that corresponded with the experimental fluxes.
To formulate the dynamic optimisation problem for minimising A sensitivity analysis was carried out by steady-state simulations
the total energy consumed in DCMD, a second outer loop was added to over a wide range of operating conditions to improve the performance
the dynamic simulation problem, as shown in Fig. 6. of the DCMD process, which could not be deduced from the lab-scale
experiments. Results showed a positive effect of the feed temperature
5. Results and discussion and velocity on mass flux and energy efficiency, due to their roles in
reducing viscosity, easing the flow, and increasing water vapour
5.1. Steady-state DCMD simulation pressure on the feed side. Accordingly, it is desirable that DCMD be
operated at the highest allowable feed temperature and velocity,
A MATLAB code was written to implement the steady-state which is well in agreement with a finding of Chen et al. [31].
simulation algorithm in Fig. 4 and to accommodate step-wise Permeate inlet temperatures appeared to pose a significant effect
sequences of the studied operating conditions that had been applied in reducing DCMD mass flux when increasing from 10 to 35 °C. Below
in the lab-scale DCMD experiments. This step was aimed at validating 10 °C, the permeate inlet temperature seemed to have minor
the semi-empirical models, developed in [18], for heat and mass influence on DCMD mass flux. This phenomenon was anticipated
transfer in the shell side of an HFM. Results of the simulated DCMD and it appeared to be consistent with the finding in a numerical study
on DCMD by Alklaibi and Lior [19]. Because it is known that decreasing
the water temperature from 10 °C to 0 °C results in an increase of only
about 10% of the process' mass driving force, it is not necessary to
maintain permeate temperature below 10 °C.
Permeate velocities had a minor role in improving the DCMD mass
flux, but a permeate velocity below 0.3 m s− 1 would cause a rapid

Fig. 6. Algorithm for dynamic optimisation of DCMD in an HFM. (G0 is a vector of


selected operating conditions subjected to optimisation; the dynamic loop refers to
Fig. 5.) Fig. 8. Comparison between simulated and experimental heat fluxes.
34 V.A. Bui et al. / Desalination 259 (2010) 29–37

Table 3
Comparison between simulated and experimental heat and mass fluxes.

Module Mass flux ratio Heat flux ratio

Jsimulated/Jexperimental R2 Qsimulated/Qexperimental R2

PV80 1.0134 ± 0.0056 0.9927 0.9800 ± 0.0062 0.9743


PV65 1.0050 ± 0.0086 0.9868 0.9890 ± 0.0069 0.9769
PV37 1.0068 ± 0.0053 0.9949 1.0003 ± 0.0061 0.9827
HL50 0.9864 ± 0.0117 0.9634 0.9761 ± 0.0089 0.9573
HL31 1.0324 ± 0.0102 0.9764 1.0194 ± 0.0102 0.9455

R2, coefficient of determination in linear regression analysis of the heat and mass flux
ratios.

decrease in mass flux. Consequently, when looking at maintaining an


adequate mass flux, it is recommended to operate the DCMD process
at a permeate velocity above 0.3 m s− 1. On the other hand, operating
DCMD at a high permeate velocity may be unbeneficial, since the
improvement in mass flux was not significant.
Fig. 10. Variation of optimal permeate inlet temperatures for maximum DCMD energy
The above recommendations were solely based on the improve- efficiency due to feed concentration. (Arrow shows feed concentration Cf increasing
ment of mass flux. For an efficient operation, DCMD energy efficiency, from 30 to 60% w/w, data simulated for module HL50 at feed inlet temperature
known as a ratio of useful heat (Qdiff) over total heat (Qm), should be Tfi = 40 °C and velocities of the two streams ωf = ωp = 0.8 m s− 1).
taken into consideration. Experiment results showed that permeate
inlet temperature had a significantly greater influence on DCMD
energy efficiency than permeate velocity as illustrated in Fig. 9. There hand, the permeate inlet temperature and velocity should be
exists a peak in energy efficiency at permeate inlet temperatures maintained at their optimal values, which varied due primarily to
within the range from 20 to 32 °C, depending primarily on the feed the feed concentration. This finding is of significant benefit in tackling
concentration, while the permeate velocity imposed only a slight the problem of high energy demand within the DCMD process,
effect. The higher the feed concentration, the lower the permeate especially for concentration of thermally sensitive liquid products
temperature required for an efficient DCMD operation as illustrated in when low feed temperature is required for preserving the product
Fig. 10. The highest energy efficiency achieved was 49.9% when feed quality.
concentration was 30%. Thus, taking into consideration both mass flux
and energy efficiency, it is highly recommended that the DCMD 5.2. Dynamic simulation and optimisation for batch DCMD operation
process be operated at a permeate inlet temperature much higher
than 10 °C, and with a velocity greater than 0.3 m s− 1, while the feed The dynamic simulation was designed to reproduce a real-time
inlet temperature and velocity should both be set at the highest batch concentration operation by the available lab-scale DCMD
allowable levels. system, at various operating conditions, in order to concentrate an
The maximum energy efficiencies found for all HFMs through the amount of 1.5 kg glucose solution from 30% to 60% w/w. To enable the
steady-state simulations were significantly higher than those dynamic simulation, the operating conditions were set to their
obtained in the lab-scale experiments where the permeate inlet respective upper and lower bounds as such to ensure that water
temperature had been maintained at 10 °C. Based on all these was always being removed from the feed into the permeate. These
revelations, it was concluded that energy efficiency of the DCMD bounds were found in the steady-state simulations, where the feed
process could be maximised if the process were operated at the concentration was Cf = 60% w/w, and the minimum mass flux of
highest allowable feed inlet temperature and velocity. On the other J = 0.1 kg m− 2 h− 1.
The dynamic simulation of DCMD at various operating conditions
on the feed side revealed that both feed inlet temperature and velocity
improved the DCMD performance by enhancing the mass flux,
reducing the operating time and energy consumption, thus improving
the energy efficiency, as illustrated in Fig. 11 for module HL50.
Operating DCMD at the feed inlet temperature Tfi = 20 °C and velocity
ωf = 0.4 m s− 1 would encounter an energy efficiency of around 17%;
meanwhile the energy efficiency could reach up to 34% if the system
was operated at Tfi = 40 °C and ωf = 0.8 m s− 1. In conclusion, a batch
DCMD should be operated at the highest allowable feed temperature
and velocity, while still maintaining the product qualities as discussed
previously.
Unlike on the feed side, there existed conditions on the permeate
side where the required total energy was minimised, hence the
process' energy efficiency was maximised, as shown in Fig. 12.
Applying the lowest temperature and highest velocity on the
permeate side could only improve the mass flux, due to increasing
transmembrane water vapour pressure gradient, but attained the
lowest energy efficiency, or the highest energy consumption.
For this reason, a dynamic optimisation problem, OPTIM, was
Fig. 9. Effect of permeate inlet temperature and velocity on DCMD energy efficiency
formulated to find optimal conditions for the permeate side only,
with feed being glucose solution. (Data simulated for module HL50 at feed inlet while maintaining those on the feed side at their highest allowable
temperature Tfi = 40 °C, velocity ωf = 0.6 m s− 1 and concentration Cf = 30% w/w.) levels. All these operating conditions were designed to remain
V.A. Bui et al. / Desalination 259 (2010) 29–37 35

Table 4
Optimal operating conditions and the corresponding energy required in batch DCMD
concentration, found by OPTIM.

Module PV80 PV65 PV37 HL50 HL31

Optimal conditions on the permeate side:


Velocity ωp (m s− 1) 0.41 0.66 0.46 0.49 0.71
Inlet temperature Tpi (°C) 24.37 23.52 22.69 26.25 26.21
3
EDCMD × 10 (kJ) 15.66 12.22 13.38 4.43 5.80
EEDCMD (%) 11.57 14.88 13.56 41.04 31.31
Processing time (min) 1939 1008 643 478 539
Calculation timea (min) 123 321 152 156 275

Under conditions in lab-scale DCMD experiments: Tpi = 10 °C, ωp = 1.2 m s− 1


EDCMD × 103 (kJ) 21.24 16.36 17.41 5.58 7.30
EEDCMD (%) 8.57 11.18 10.48 32.76 25.02
Processing time (min) 1243 718 395 270 306
Total energy saving (%) 26.3 25.3 23.1 20.6 20.5

EDCMD and EEDCMD, total energy and total energy efficiency of the whole concentration
operation, estimated by Eqs. (15) and (17) respectively; data obtained at feed inlet
temperature Tfi = 40 °C and velocity ωf = 0.6 m s− 1 by simulation.
a
Calculation time of OPTIM programme performed by a desktop using Intel Pentium
Fig. 11. Effect of feed inlet temperature and velocity on total energy efficiency in batch
Core 2 Duo 1.8 GHz.
DCMD concentration. (Data simulated for module HL50 at permeate inlet temperature
Tpi = 10 °C and velocity ωp = 1.2 m.s− 1, total energy efficiency by Eq. (17).)

unchanged throughout the entire operation. The following were the than that at conditions presented in the lab-scale DCMD experiments.
inputs and limits applied in the programme: An option to speed up the process is to increase the membrane area in
a practical application.
• Feed solution: initial weight M0 = 1.5 kg and concentration C0 = 30%
w/w.
5.3. Testing optimisation results
• Conditions on the feed side: Tfi = 40 °C, ωf = 0.6 m s− 1.
• Bounds on the permeate side: 10 °Cb Tpi b 30 °C and 0.3 b ωp b 1.3 m s− 1
In OPTIM, temperatures and velocities of the two streams were
• Constraints imposed: the minimum mass flux, J ≥ 0.5 kg m− 2 h− 1,
designed to remain unchanged throughout the entire operation. To test
and the maximum pressure drop through the module on the
the results by OPTIM, the lab-scale DCMD system described in Fig. 3 was
permeate side, Δpp ≤ 150 kPa.
used to concentrate 1.5 kg of glucose solution from 30% to 60% w/w. The
It should be noted that the mass flux constraint could be activated test, referred to as Test1, was conducted on the best performing module
only at the end of the operation, when feed concentration reached 60% HL50 only, under the operating conditions listed in Table 4.
w/w. The found optimal conditions and the corresponding total It should be noted that the inlet temperatures of the two streams
energy required for all modules are listed in Table 4. fluctuated over the processing time, although originally expected to
It can be seen from these results that the optimal permeate inlet remain constant, and therefore these temperature changes were also
temperature and velocity fell respectively within a moderate range of recorded, as were temperature changes from the inlet to the outlet of
between 22 and 26 °C, and from 0.4 to 0.7 m s− 1. Operating DCMD at the corresponding feed and permeate streams. These temperature
these optimal conditions could result in a major reduction of up to changes and the pressure drops through the HFM by the two streams
26.3% of the total energy required for module PV80. However, the were then used in Eqs. (15) and (16), to calculate the total energy
total energy saved was only 20.6% for the best available module HL50. consumed by the DCMD process at a time interval of 15 min. The
As a trade-off, saving energy requires a much longer processing time measured and simulated mass flux and feed concentration changes in
the concentration operation are shown in Fig. 13.
The experimentally obtained mass flux appeared to be up to 15%
higher than the simulated mass flux over the first 2.5 h, leading to a
faster increase in feed concentration and a shorter processing time
compared to those determined by simulation under the same
conditions. The differences between the simulated and measured
mass fluxes over the first 4 h, shown in Fig. 13, may have been
dominated by the fluctuations of the controlled operating conditions
in Test1 in the regions that favoured the DCMD process performance.
These differences were acceptable since the flow rates of the two
streams and the feed inlet temperature were always maintained at or
above the desired levels, while the permeate inlet temperature always
at or below 26.2 °C. Over the last 2 h, these differences seemed to be
dominated by the higher feed concentration, which was accumulated
over the previous first four hours, in Test1 than in the simulation. It
might also be speculated that these differences have lain within the
uncertainty of the models in Eqs. (10) and (11).
Similar plots were also obtained in another batch DCMD operation,
referred to as Test2, which differed from Test1 in the conditions that
were applied on the permeate side. These included the permeate inlet
Fig. 12. Effect of permeate inlet temperature and velocity on total energy efficiency in
batch DCMD concentration. (Data simulated for module HL50 at feed inlet temperature
temperature Tpi = 10 °C and velocity ωp = 1.2 m s− 1, which were
Tfi = 40 °C and velocity ωf = 0.6 m.s− 1, total energy efficiency by Eq. (17), lines are similar to those applied in the DCMD lab-scale experiments. This test
contours.) was designed to confirm the performance of the dynamic simulation
36 V.A. Bui et al. / Desalination 259 (2010) 29–37

food industry. As the optimisation carried out in this work focussed


solely on reducing the energy consumption, a full scale optimisation
of DCMD, including other costs such as labour, plant site and
amortisation of equipment, is still required for further investigation.
With an achievable total energy efficiency of up to only 41.04% for
the best available module HL50 when operated at the optimal
conditions (as found by OPTIM), DCMD would still face a serious
challenge in energy consumption. Therefore, a method for reducing
the energy consumption in a DCMD concentration operation at low
temperatures is still required to ensure the survival of this technology.

Nomenclature
Capital letters
Ai, Ao internal, external surface area of hollow fibres, (m2)
Alm logarithmic-mean area of the insulation for an HFM, (m2)
Fig. 13. Changes of mass flux and feed concentration in batch DCMD concentration
under the optimal conditions found by OPTIM (Test1). (Cexp, measured feed C concentration of a solution, (% w/w)
concentration in °Brix, Csim, simulated feed concentration in % w/w, Jexp and Jsim, Cpx specific heat capacity of material in the (x) domain,
measured and simulated mass flux; data were obtained for module HL50 operating at (J kg− 1 K− 1)
feed inlet temperature Tfi = 40 °C and velocities ωf = 0.6 m s− 1, permeate inlet J mass flux, (kg m− 2 s− 1)
temperature Tpi = 26.2 °C and velocity ωp = 0.5 m s− 1.)
Kf convective mass transfer coefficient on the feed side, (m s− 1)
Km membrane's vapour permeability, (kg m− 2 s− 1 Pa− 1)
programme, and to show if OPTIM could produce any significant L length of hollow fibre, (m)
energy savings. Results of Test1 and Test2 are summarised in Table 5. Px water vapour pressure, (Pa)
With differences in the processing times and total energy obtained by Qx heat flux gained or lost by the (x) stream (x = F for feed or
the experiments and by simulations respectively being only 5.8% and x = P for permeate), (W)
1.8%, it was concluded that the dynamic simulation programme could Qloss heat lost through the membrane module insulation, (W)
reproduce the DCMD batch concentration operation fairly well. Qx heat flux across the (x) domain, (W)
Comparison between the total energy consumed in the two tests Qdiff heat flux due to mass diffusion, (W)
showed that operating DCMD under conditions gathered from OPTIM Qcond heat flux due to conduction, (W)
could save about 23% of total energy, which is a significant saving in
the DCMD process.
Lower case letters
6. Conclusions dhx equivalent diameter of flow channel x (x = f or p), (m)
hx heat transfer coefficient at the x-domain (x = f, p or m),
A computational framework has been developed to simulate and (W m− 2 K− 1)
optimise the DCMD process in a counter-current flow, hollow fibre kx thermal conductivity, (W m− 1 K− 1)
configuration. The framework was transformed accordingly into ṁ mass flow rate of a stream, (kg s− 1)
MATLAB programmes, to simulate the static and dynamic states of ri / ro internal/external radius of hollow fibres, (m)
mass and heat transfer in a real-time DCMD batch concentration
operation.
Steady-state simulation suggested that DCMD could be operated Greek letters
most efficiently at the highest allowable feed temperature and ΔHv latent heat of vaporisation of water, (J kg− 1)
velocity, while those on the permeate side not below 10 °C and ΔPx vapour pressure difference across x-domain
0.3 m s− 1, respectively. α, β, φ exponents in semi-empirical models
Dynamic optimisation, OPTIM, was carried out for a batch DCMD µ viscosity, (Pa s)
operation. Operating the DCMD process under the optimal conditions ω fluid velocity, (m s− 1)
could result in a reduction in energy consumption of up to 26.3%,
thereby assisting to tackle the energy challenge faced by DCMD when
operated at low feed temperatures, and fostering its application in the Subscripts
amb ambient
b in the bulk
Table 5
Comparison between experiments and simulations in a batch DCMD operation
f feed or feed domain
concentrating 1.5 kg glucose solution from 30 to 60% w/w. fm at feed membrane surface
fi, fo feed inlet, outlet
Operating conditions Test 1 (by OPTIM) Test 2
m membrane, or membrane surface
on the permeate side
Tpi =26.2 °C, ωp =0.5 m s−1 Tpi =10 °C, ωp =1.2 m s−1 p permeate or permeate domain
Experiment Simulation Experiment Simulation pl membrane polymer
Processing time (min) 450 478 255 270
pm at permeate membrane surface
EDCMD × 103 (kJ) 4.27 4.35 5.54 5.58 pi, po permeate inlet, outlet
Final concentration 60.2°Brix 60.1% w/w 60.0°Brix 60.2% w/w

Comparison between experiment and simulation


Processing time (%) − 5.8% − 5.6% Dimensionless numbers used in semi-empirical equations
EDCMD (%) − 1.8% − 0.72% Nu Nusselt number
EDCMD, total energy consumption; data obtained for module HL50 at feed inlet
Gz Graetz number
temperature Tfi = 40 °C and velocity ωf = 0.6 m s− 1; comparison percentage showing Pr Prandtl number
the respective data by experiment relative to that by simulation. Re Reynolds number
V.A. Bui et al. / Desalination 259 (2010) 29–37 37

Sh Sherwood number [14] M. Khayet, M.P. Godino, J.I. Mengual, Possibility of nuclear desalination through
various membrane distillation configurations: a comparative study, Int. J. Nucl.
Sc Schmidt number Desalin. 1 (1) (2003) 30–46.
[15] M.N. Ozisik, Heat transfer — a basic approach, Mech. Eng. Series, McGraw-Hill
Intnl Editions, Singapore, 1985.
Acknowledgements [16] F.M. White, Heat and Mass Transfer, Addison-Wesley Publishing Company, Inc.,
USA, 1991.
The authors would like to thank the University of Western Sydney [17] W.L. McCabe, J.C. Smith, P. Harriott, Unit operations of chemical engineering,
Chem. Eng. Series4th ed., McGraw-Hill Intnl Editions, 1985.
for the International Postgraduate Research Scholarship for Viet Bui, [18] V.A. Bui, L.T.T. Vu, M.H. Nguyen, Modelling the simultaneous heat and mass
Siemens Water Technology-Australia for supporting the project with transfer of direct contact membrane distillation in hollow fibre modules, J. Membr.
hollow fibres, and Dr. J. Muller for his technical advice. Sci. 353 (1–2) (2010) 85–93.
[19] A.M. Alklaibi, N. Lior, Comparative study of direct-contact and air-gap membrane
distillation processes, Ind. Eng. Chem. Res. 46 (2) (2007) 584–590.
References [20] S. Ma, L. Song, S.L. Ong, W.J. Ng, A 2-D streamline upwind Petrov/Galerkin finite
element model for concentration polarization in spiral wound reverse osmosis
[1] R.W. Schofield, A.G. Fane, C.J.D. Fell, Heat and mass transfer in membrane modules, J. Membr. Sci. 244 (1–2) (2004) 129–139.
distillation, J. Membr. Sci. 33 (3) (1987) 299–313. [21] A. Subramani, S. Kim, E.M.V. Hoek, Pressure, flow, and concentration profiles in
[2] M.S. El-Bourawi, Z. Ding, R. Ma, M. Khayet, A framework for better understanding open and spacer-filled membrane channels, J. Membr. Sci. 277 (1–2) (2006) 7–17.
membrane distillation separation process, J. Membr. Sci. 285 (1–2) (2006) 4–29. [22] M. Khayet, M. Godino, J. Mengual, Study of asymmetric polarization in direct
[3] A.K. Manna, M. Sen, A.R. Martin, P. Pal, Removal of arsenic from contaminated contact membrane distillation, Sep. Sci. Technol. 39 (1) (2004) 125–147.
groundwater by solar-driven membrane distillation, Environ. Pollut. 158 (3) [23] L. Martínez-Díez, F.J. Florido-Díaz, M.I. Vázquez-González, Study of polarisation
(2010) 805–811. phenomena in membrane distillation of aqueous salt solutions, Sep. Sci. Technol.
[4] D. Hou, J. Wang, X. Sun, Z. Luan, C. Zhao, X. Ren, Boron removal from aqueous 35 (10) (2000) 1485–1501.
solution by direct contact membrane distillation, J. Hazard. Mater. 177 (1–3) [24] V.A. Bui, M.H. Nguyen, J. Muller, Characterisation of the polarisations in osmotic
(2010) 613–619. distillation of glucose solutions in hollow fibre module, J. Food Eng. 68 (3) (2005)
[5] A.M. Alklaibi, N. Lior, Membrane-distillation desalination: status and potential, 391–402.
Desalination 171 (2) (2004) 111–131. [25] J.I. Mengual, et al., Osmotic distillation through porous hydrophobic membranes,
[6] K.W. Lawson, D.R. Lloyd, Membrane distillation, J. Membr. Sci. 124 (1) (1997) J. Membr. Sci. 82 (1–2) (1993) 129–140.
1–25. [26] R. Thanedgunbaworna, R. Jiraratananona, M.H. Nguyen, Shell-side mass transfer
[7] F. Lagana, G. Barbieri, E. Drioli, Direct contact membrane distillation: modelling of hollow fibre modules in osmotic distillation process, J. Membr. Sci. 290 (1–2)
and concentration experiments, J. Membr. Sci. 166 (1) (2000) 1–11. (2007) 105–113.
[8] V.A. Bui, M.H. Nguyen, J. Muller, Challenge and opportunity of direct contact [27] C. Fernandez-Pineda, M.A. Izquierdo-Gil, M.C. Garcia-Payo, Gas permeation and
membrane distillation in liquid food concentration, in: B.R. Young, D.A. Patterson, direct contact membrane distillation experiments and their analysis using
X.D. Chen (Eds.), Chemeca 2006, Engineers Australia, Auckland, New Zealand, different models, J. Membr. Sci. 198 (1) (2002) 33–49.
Sept 17–21st 2006, Article 137. [28] L. Martinez, F.J. Florido-Diaz, Theoretical and experimental studies on desalination
[9] E. Curcio, G. Barbieri, E. Drioli, Operazioni di distillazione a membrana nella using membrane distillation, Desalination 139 (1–3) (2001) 373–379.
concentrazione dei succhi di fruitta, Ind. Bevande (April 29 2000) 113–120. [29] R.W. Schofield, A.G. Fane, C.J.D. Fell, Gas and vapour transport through
[10] V.D. Alves, I.M. Coelhoso, Orange juice concentration by osmotic evaporation and microporous membranes. II. Membrane distillation, J. Membr. Sci. 53 (1–2)
membrane distillation: a comparative study, J. Food Eng. 74 (1) (2006) 125–133. (1990) 173–185.
[11] V.A. Bui, M.H. Nguyen, The role of operating conditions in osmotic distillation and [30] A.M. Alklaibi, N. Lior, Heat and mass transfer resistance analysis of membrane
direct contact membrane distillation — a comparative study, Int. J. Food Eng. 2 (5) distillation, J. Membr. Sci. 282 (1–2) (2006) 362–369.
(2006) Article 1. [31] T.C. Chen, C.D. Ho, H.M. Yeh, Theoretical modelling and experimental analysis of
[12] L. Martinez-Diez, F.J. Florido-Diaz, M.I. Vazquez-Gonzalez, Study of evaporation direct contact membrane distillation, J. Membr. Sci. 330 (2009) 279–287.
efficiency in membrane distillation, Desalination 126 (1–3) (1999) 193–198.
[13] V.A. Bui, M.H. Nguyen, J. Muller, The energy challenge of direct contact membrane
distillation in low temperature concentration, Asia-Pacific J. Chem. Eng. 2 (5)
(2007) 400–406.

You might also like