Sei sulla pagina 1di 26

Journal of Membrane Science 285 (2006) 429

Review

A framework for better understanding membrane


distillation separation process
M.S. El-Bourawi a , Z. Ding a , R. Ma a, , M. Khayet b,1
a

State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing 100029, PR China
Department of Applied Physics I, Faculty of Physics, University Complutense of Madrid, Avda. Complutense of Madrid s/n, 28040 Madrid, Spain
Received 14 March 2006; received in revised form 29 July 2006; accepted 1 August 2006
Available online 8 August 2006

Abstract
Membrane distillation (MD) is an emerging technology for separations that are traditionally accomplished by conventional separation processes
such as distillation or reverse osmosis. Since its appearance in the late of the 1960s and its development in the early of 1980s with the growth
of membrane engineering, MD claims to be a cost effective separation process that can utilize low-grade waste and/or alternative energy sources
such as solar and geothermal energy. As an attractive separation process, MD has been the subject of worldwide academic studies by many
experimentalist and theoreticians. Unfortunately from the commercial stand point, MD has gained only little acceptance and yet to be implemented
in industry. The major barriers include MD membrane and module design, membrane pore wetting, low permeate flow rate and flux decay as
well as uncertain energetic and economic costs. This study is an attempt to establish a framework for better understanding the MD process and to
consider all possible solutions developed so far to overcome its barriers. Unlike the usual trend pursued in review papers, MD studies have been
cited in the present manuscript and classified in tables according to their most important contribution in MD development. These tables cover most
important aspects of the MD process and are presented in a simple manner for a glance understanding the effects of different factors and operating
variables on the productivity of each MD configuration. Among the different MD papers, those involving theoretical models are pointed out. The
areas within the MD field that are either usually or rarely studied are highlighted. Some useful technical discussions based on acquired knowledge
from experience and information gathered from MD literature are included. In some way, this paper will help new researchers in the field of MD
to quickly be updated avoiding repetition of already known studies. In fact, although the effects of some operating parameters are generally agreed
upon, still new researches appear with almost the same results.
2006 Published by Elsevier B.V.
Keywords: Membrane distillation; Different MD configurations; Modeling; Factors affecting MD; Long-term MD performance

Contents
1.
2.
3.
4.
5.
6.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
MD configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Membrane distillation application areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Approaches followed for MD improvements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Modeling in MD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Operating variables affecting MD process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1. Feed temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2. Feed inlet concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.3. Feed circulation velocity and stirring rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.4. Permeate inlet temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.5. Temperature difference and mean temperature effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author. Fax: +86 10 64423610.


E-mail addresses: r.ma@mail.buct.edu.cn (R. Ma), khayetm@fis.ucm.es (M. Khayet).
Tel.: +34 91 3944454; fax: +34 91 3945191.

0376-7388/$ see front matter 2006 Published by Elsevier B.V.


doi:10.1016/j.memsci.2006.08.002

5
7
8
8
10
11
11
11
13
13
13

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

7.

8.

9.

6.6. Permeate flow velocity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


6.7. Vapor pressure difference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Membrane parameters affecting MD process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.1. Membrane thickness. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2. Membrane porosity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3. Membrane pore size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.4. Pore size distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.5. Pore tortuosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.6. Membrane surface chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.7. Membrane module geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Framework for better understanding MD process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.1. Flow chart towards the industrialization of MD process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.2. The different aspects of the MD process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.2.1. Heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.2.2. Mass transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.3. Rate controlling step in MD process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.4. Membrane fouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.4.1. Crystallization fouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.4.2. Biological fouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.4.3. Particulate and corrosion fouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.5. Long-term MD performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.6. Energy and maintenance cost evaluation in MD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Membrane distillation (MD) is a thermally driven process,
in which only vapor molecules are transported through porous
hydrophobic membranes. The liquid feed to be treated by MD
must be in direct contact with one side of the membrane and
does not penetrate inside the dry pores of the membranes. The
hydrophobic nature of the membrane prevents liquid solutions
from entering its pores due to the surface tension forces. As a
result, liquid/vapor interfaces are formed at the entrances of the
membrane pores. The MD driving force is the transmembrane
vapor pressure difference that may be maintained with one of the
four following possibilities (see Fig. 1) applied in the permeate
side [17]:
(i) An aqueous solution colder than the feed solution is maintained in direct contact with the permeate side of the membrane giving rise to the configuration known as direct contact membrane distillation (DCMD). The transmembrane
temperature difference induces a vapor pressure difference.

14
14
14
14
15
15
16
16
16
16
17
17
17
17
21
22
23
23
23
23
24
24
25
25
26

Consequently, volatile molecules evaporate at the hot liquid/vapor interface, cross the membrane in vapor phase and
condense in the cold liquid/vapor interface inside the membrane module.
(ii) A stagnant air gap is interposed between the membrane and
a condensation surface. In this case, the evaporated volatile
molecules cross both the membrane pores and the air gap to
finally condense over a cold surface inside the membrane
module. This MD configuration is called air gap membrane
distillation (AGMD).
(iii) A cold inert gas sweeps the permeate side of the membrane carrying the vapor molecules and condensation takes
place outside the membrane module. This type of configuration is termed sweeping gas membrane distillation
(SGMD).
(iv) Vacuum is applied in the permeate side of the membrane
module by means of a vacuum pump. The applied vacuum
pressure is lower than the saturation pressure of volatile
molecules to be separated from the feed solution. In this
case, condensation occurs outside of the membrane mod-

Fig. 1. Different types of MD configurations.

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

ule. This MD configuration is termed vacuum membrane


distillation (VMD).
Each one of the above possibilities has its advantages and
inconveniencies for a given application. The potential advantages of MD process in comparison to the conventional separation processes rely on the lower operating temperature and
hydrostatic pressure. Feed solutions having temperatures much
lower than its boiling point under pressures near atmosphere can
be used.
It must be pointed out that in MD, the membrane itself
acts only as a barrier to hold the liquid/vapor interfaces at
the entrance of the pores and it is not necessary to be selective as required in other membrane processes such as pervaporation. The main requirements for MD process are that
the membrane must not be wetted and only vapor and noncondensable gases are present within its pores. The pores size
of the membranes used in MD lies between 10 nm and 1 m.
To avoid pore wettability, the membrane material must be
hydrophobic with high water contact angle and small maximum pore size. Hydrophobic microporous membranes such
as those made from polypropylene (PP), polyethylene (PE),
polytetrafluoroethylene (PTFE), and polyvinylidene fluoride
(PVDF) meets these requirements. However pore wettability
may occur and permeate quality may be affected if solutions
with surface active components are brought in direct contact
with the membrane surface or if the transmembrane hydrostatic pressure exceeds the liquid entry pressure (LEP), which
is characteristic of each membrane. MD is therefore, a process
mainly suited for applications in which the major component is
water.
MD is first patented in 1963 [8] and the first MD paper published in Journal appeared 4 years later, in 1967 [9]. At that time,
interest in this process has been faded quickly loosing its brightness due partly to the observed lower MD production compared
to the reverse osmosis technique. MD process has recovered
interest within the academic communities in the early of 1980s
when novel membranes and modules with better characteristics
became available [1012]. This alongside with the actual merits
of MD capability to utilize low-grade waste and/or alternative
energy sources, such as solar and geothermal energy, made it
more promising separation technique [13,14]. As a result of
intensive worldwide studies by many experimentalists and theoreticians investigators, advancements of MD process lead to the
development of the previously cited MD configurations. Nowadays, the importance and realization of MD process as a useful
separation technique can be detected from the increasing number of published papers available in membrane literature. Fig. 2
illustrates the time frame of the works (those we were able to
detect) performed on MD since its first appearance. Although the
1980s showed the recover interest in MD, the real vast growing
interest in MD noticeably started from the late 1990s. During the
time period from 1999 to 2005, for instance, the number of published studies on MD is much more than that performed during
the period from 1963 to 1998. Nonetheless, from a commercial
standpoint, MD is not implemented yet in industry. The main
barriers to commercial implementation of MD and preventing

Fig. 2. Time frame of the works performed on MD process.

MD from being a viable separation technology include:


(i) Relatively low permeate flux compared to other separation
techniques such as reverse osmosis.
(ii) Permeate flux decay due to concentration and temperature
polarization effects, membrane fouling and total or partial
pore wetting.
(iii) Membrane and module design for MD.
(iv) High Thermal energy consumption: uncertain energy and
economic costs for each MD configuration and application.
In attempts to overcome these challenges, investigators
launched all possible efforts to prove MD validity as a successful
separation technique. Benefiting from the low temperature and
transmembrane hydrostatic pressure required to perform MD
operations, several approaches to make the MD as a viable separation technique were proposed. These approaches ranged from
finding new areas of MD applicability and the cooperation of
MD with other processes as a pre-treatment or post-treatment
step, to researches devoted in preparation of membranes together
with MD modules and studies of factors affecting MD production associated to the application of some enhancement techniques.
Although the effect of most of the design factors and operating variables are well understood and documented in the MD
literature, scattered or inconsistency information exists about
some others. Some of the recognized MD configurations are
usually investigated, while others are rarely studied and far from
being fully controlled. Attempts of providing successful mathematical models that can describe the transport process for each
MD configuration are widely available in MD literature. Despite
of their precision in estimating effects of different parameters
on permeate fluxes, those models are rarely used by others, as
they are based on special experimental facilities that usually
differ from one to another investigator. Most of the models are
based on at least one adjustment parameter frequently taken as
pore tortuosity factor. Other semi-empirical models are based on
the applicability of the heat and mass transfer correlations that
are originally developed for rigid nonporous heat exchangers
and recently questioned by some authors [15,16]. Furthermore,
design of membranes and modules for MD applications is still
lacking in spite of the fact that very few papers in this subject
appear from time to time [1,1719]. It is worth quoting that,
although the effect of some factors and variables are generally

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

agreed upon, still some papers are published reporting almost


the same results and conclusions. For example, the effects on the
MD flux of the operating variables such as feed and permeate
temperatures and flow velocities have been widely investigated
experimentally and corroborated with different theoretical models, but still some published studies on the same topics appear
in various journals repeating the investigation of these effects.
Meanwhile, membrane factors such as membrane pore size distribution and pore tortuosity, membrane fouling, energy analysis
and heat recovery are rarely studied or even ignored despite
of the dominant role that they may play in the future of MD
process. On the other hand, within the MD literature one can
easily observe that among the MD configurations, the DCMD
mode is the most studied. This may be related to its ease of
handling as the condensation step is carried out inside the membrane module. On the contrary, VMD and SGMD are the least
studied configurations. This may be due to the fact that the last
two configurations require external condensers, applied outside
the membrane module, resulting in a more complicated MD
system. Therefore, as this paper aims to, there is a need to
gather all possible available information and results concerning the MD process in general, each of its configurations and
the effect of operating variables and factors that affect their
performances. This fact may lead to the development of a framework for better understanding the MD process and highlight
areas that still need to be investigated further in order to avoid
repeatability.
2. MD congurations
It is reported in membrane literature [1,4,16,20,21] that
DCMD is the most studied MD configuration as stated in the
previous section. This is due to the fact that condensation step
is carried out inside the membrane module leading in this way
to a simple operation mode without the need of external condensers like in SGMD and VMD configurations. Fig. 3 shows a
comparison between the numbers of gathered papers published
in journals for each MD configuration together with the corresponding number of papers involving theoretical models. It can
be seen that among the published studies, more than 60% are
focused on the DCMD configuration in spite of the fact that the
heat loss by conduction through the membrane matrix in DCMD,
which is considered heat loss in MD, is higher than that of the

Fig. 3. Number of papers published in refereed journals for each MD configuration and corresponding papers involving theoretical models.

other MD configurations. On the other hand, VMD configuration exhibiting higher permeate flux and practically negligible
heat transfer by conduction through the membrane is the least
used configuration together with the SGMD one [21,22]. In fact,
in Fig. 3, it can be observed that less than 15% of the relevant
references deal with VMD configuration. It is worth quoting
that in VMD, the transmembrane hydrostatic pressure must be
kept below the minimum feed liquid entry pressure (LEP) of the
membrane. Because in this configuration vacuum is applied in
the permeate side and the permeate flux is higher for lower permeate pressure, the risk of membrane pore wetting is very high
[5,23]. The Laplace (Cantor) equation provides the relationship
between the membranes largest allowable pore size (rmax ) and
the related operating conditions [1].
LEP > Pinterface = Pliquid Pvapor =

2BL cos
rmax

(1)

where B is a geometric factor determined by pore structure, L


the liquid surface tension and is the liquid/solid contact angle.
More details are provided in Ref. [1]. It is worth quoting that
membranes of smaller pore sizes (i.e. less than 0.45 m) than
the ones used in the other MD configurations are used in VMD.
It must be stated that the AGMD configuration exhibits the
lower permeate flux compared to that of the other MD configurations. On the other hand, the introduced air gap in the
membrane module reduces considerably the heat loss by conduction and the temperature polarization, which are higher in
DCMD configuration. However, AGMD configuration presents
a new resistance to heat and mass transfer due to the presence
of the air gap between the membrane and the condensation surface, which results in lower permeate fluxes. The influence of
this new resistance on the transmembrane flux and on the process
selectivity depends on the effective air gap width [5,24,25].
The SGMD was developed to provide an intermediate
solution between the DCMD and the AGMD configurations
[4,2628]. In SGMD, there is a combination of both the low
conductive heat loss of AGMD process with the reduced mass
transfer resistance of DCMD. As in AGMD, there is a gas barrier between the membrane and the cold surface, which results
in reduction in heat loss. However the gas in SGMD is not
stationary and sweeps the membrane, which in turn results in
enhancing the mass transfer coefficient and leading to higher
permeate flux than in AGMD. In addition, the SGMD configuration exhibits higher permeate flux and evaporation efficiency
than the DCMD process [22,2628]. Furthermore, the question
of energy efficiency is very important in MD. In fact, in this
process heat recovery can be carried out by an external heat
exchanger or by utilizing an internal heat recovery function
where the inlet feed liquid is circulating through a condenser
channel to warm up [14]. It is worth mentioning here that heat
recovery in SGMD configuration is quite difficult. A gas source
is required for SGMD permeate production and the costs associated to the transported gas are very high. These facts may also
explain why this configuration has received only little attention
in comparison to DCMD configuration.
Generally, there are three major energetic inefficiencies in
MD. The first is due to temperature and concentration polar-

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

ization effects, the second is due to air/gas trapped within the


membrane pores which impose a resistance to mass transfer,
while the third is due to the heat loss by conduction through
the membrane matrix. It must be pointed out that the thermal
energy consumption in the non-isothermal MD process must be
evaluated adequately for each MD configuration. Both the MD
system and the operation costs are needed to be determined.
The high thermal energy needed can be overcame by efficient
heat recovery and/or by using low energy-sources or waste heat,
which may be obtained for example from nuclear power stations
[22,29].
3. Membrane distillation application areas
Table 1 broadly summarizes areas where the different MD
configurations were examined, on laboratory scale, and found
to be applicable. In more details, MD has been applied for
separation of non-volatile components from water like ions, colloids, macromolecules [1,2,23,2832], for the removal of trace
volatile organic compounds from water such as benzene, chloroform, trichloroethylene [1,3338] or the extraction of other
organic compounds such as alcohols from dilute aqueous solutions [1,3942]. MD is suited for both distilled water production
or for the concentration of aqueous solutions.
MD has been applied for water desalination, environmental
waste clean up, water reuse and food processing among others like milk and juice concentration, biomedical applications
such as water removal from blood and treatment of protein
solutions [1,2,5,24,28,3032,4351]. Separation of azeotropic
aqueous mixtures such as alcoholwater mixtures, concentration
of radioactive solutions and application for nuclear desalination,
waste water treatment in which a less hazardous waste can be
discharged to the environment specially in textile waste treatment that is contaminated with dyes, concentration of coolant
(glycol) aqueous solutions, treatment of humic acid solutions,
pharmaceutical waste water treatment and in areas where high
temperature applications lead to degradation of process fluids,
can be attractive [1,29,3942,5262]. It must be pointed out that
desalination is the most known DCMD application as near 100%
rejection of non-volatile ionic solutes is easily achieved [1,22].

In addition, due to the chemical stability of the employed membranes, MD can also be applied for the concentration of acids
[5860]. In fact, the concentration sulfuric acid, hydrochloric
acid and nitric acid by both DCMD and VMD up to 60% have
been achieved.
One of the advantages of MD over the pressure-driven processes such as RO is its ability to treat aqueous feed solutions
of very high non-volatile solute concentrations. MD is even
capable of recovering crystal valuable products from effluents.
With regards to this aspect, membrane distillation crystallization (MDC) is a relatively new attractive process that recently
has been applied to recover valuable salts in crystal form from
its effluents and the processing of brine from desalination operations [54,6366]. MDC combines both a MD module for pure
water production and a crystallizer into which the brine of the
MD unit at its supersaturation state is charged and solids are
recovered in crystal forms. As a simple example, narrow anhydrous sulfate crystal size distribution with an average size of
6070 m was produced in Ref. [64].
It is worth quoting that when volatile organic compounds are
present in the feed aqueous solutions, care must be taken in order
to avoid membrane pore wetting. In this case, very low solute
concentrations are used and the liquid entry pressure of the feed
solution into the pores is necessary to be measured [67].
4. Approaches followed for MD improvements
Various experimental methods have been applied for
improvements of MD process. The first attempt was claimed
in 1967 [68]. A multi-stage MD system capable of reusing the
latent heat of vaporization many times has been proposed. The
permeate of the first stage of MD process can be used to heat
the feed of a second stage and the permeate of the second stage
is applied to heat the feed of the third and so on. Years later,
Findley concluded in his published paper [9] that if in MD
process high temperature is applied, low cost system and longlife membranes with adequate characteristics can be obtained,
MD could be an economical method for desalination of seawater. Since then, numerous studies have been performed to
investigate the effect on MD permeate flux enhancement and

Table 1
Summary of the areas where MD process were successfully applied on laboratory scale
Application area

Desalination and pure water production from brackish water


Nuclear industry (concentration of radioactive solutions and
wastewater treatments; pure water production)
Textile industry (removal of dyes and wastewater treatment)
Chemical industry (concentration of acids, removal of VOCs from
water, separation of azeotropic aqueous mixtures such as
alcohol/water mixtures and crystallization)
Pharmaceutical and biomedical industries (removal of water from
blood and protein solutions, wastewater treatment)
Food industry (concentration of juices and milk processing) and in
areas where high temperature applications lead to degradation of
process fluids

MD configuration
DCMD

AGMD

SGMD

VMD

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

selectivity of different operating variables such as temperature,


flow rate, vacuum pressure, etc. The important effects of temperature polarization and concentration polarization on permeate
flux have been investigated [1,3]. In fact, the low MD permeate
flux is attributed mainly to the temperature polarization and in
some cases, for example in DCMD process, to the heat transfer by conduction through the membrane material [3,21,6972].
Meanwhile, concentration polarization decreases the permeate
flux and increases the risk of membrane pore wetting [71,72].
Attempts have been made to diminish these two polarization
effects by utilizing different enhancement techniques including the improvement of the channel design for flow passage
through the membrane module, the adequate membrane arrangement in the membrane modules to achieve better mixing conditions and the inclusion of turbulence promoters in the flow
channels and the use of ultrasonic systems [7375]. Furthermore, studies on permeate flux enhancement by using spacer
filled channels have been also carried out in MD [76,77]. Similarly, the effect of air and non-condensable gases entrapped
within the membrane pores in DCMD configuration has been
studied [72,7880]. Additionally, influence of membrane characteristics such as membrane thickness, porosity, mean pores
size and pore size distribution on MD flux and selectivity has
been investigated [47,8084]. It is speculated that, as occurred
in other pressure-driven membrane separation processes, membrane roughness may influence the polarization effect in MD
[81].
It is generally acknowledged that the used membranes in
most of the MD researches are actually made for other process
(i.e. microfiltration) rather than MD. Despite that most of the
required features to MD membranes are met with these commercially available membranes, the need for new membranes
fabricated especially for MD purposes have been also widely
accepted by MD investigators. However, only few investigators
have worked on the design of MD membranes [1719,8588].
With the recent advancement in this field, new promising generation of membranes has appeared that not only provides higher
permeability, but also can increase transmembrane pressure difference through minimizing the heat loss by conduction through
the non-porous portion (i.e. matrix) of the membrane. Within
this area, porous homogenous and porous composite hydrophobic/hydrophilic membranes have been made and tested for MD
applications [89,90]. On the other hand, advancements in module design, let to the emerge of new modules for the different MD
configurations, that can provide improved fluid flow dynamics
and can yield high reasonable transmembrane fluxes, through
minimizing boundary layer heat and mass transfer resistances
[23,75,9196]. Nevertheless, more must be done before getting
adequate membrane for MD applications and MD configurations
as well as appropriate MD modules [89,93].
Another approach of making MD a competitive and a viable
process is to find new areas where MD can be successfully
applied. The increasing academic interests towards MD process have lead to the recognition of the wide variety of areas for
which MD are applicable such as areas where high temperature
can not be applied and lead to degradation and denaturalization
of the involved solutes present in the feed solutions.

Cooperation of MD with other processes can extend the separation efficiency of the whole integrated system. Combination
of MD with other separation processes such as reverse osmosis
(RO), ultrafiltration (UF), nanofiltration (NF), osmotic distillation (OD) or multi-effect distillers is another area where the
benefits of MD make the whole process attractive [65,97102].
It has been reported in Ref. [65] that the introduction of a MD
stage operating with the RO brine allowed a recovery factor of
about 88%, in comparison with that of 40% obtained with the RO
unit alone. Three different membranes units, namely, NF unit for
pretreatment, a RO unit and a membrane crystallizer were also
simultaneously coupled to MD [65]. The integrated system has
resulted in a 100% water recovery and in the elimination of the
brine disposal problem, as pure crystals were produced and may
be considered as a valuable product. On the other hand, combinations between different membrane separation units for the
highest water quality production from potable water were also
tested in Ref. [98]. Units such as UF, NF, RO and MD were tested
and the highest water quality was obtained with the integrated
ROMD installation. It must be mentioned that direct treatment
of potable water in MD unit has resulted in a rapid decrease of
flux due to scale formation (fouling) on the membrane surface
[98]. A combination of MD and OD processes in a single system,
termed membrane osmotic distillation (MOD), to obtain higher
water fluxes and heat efficiencies was proposed in Ref. [99] with
brine circulated in the cold side. Based on their experimental
investigation, they found that MOD has the advantage of achieving higher water flux and heat efficiency simultaneously. In a
recent study, incorporation of MD to enhance the flux and solute
rejection in an innovation system that combines three different
membrane processes for wastewater reclamation in space, was
investigated [100]. The originally innovated system consisted
of RO elements, a direct osmosis (DO) pretreatment process
and a combined direct osmotic/osmotic distillation (DO/OD)
pretreatment process. In this case, MD acts as a possible solution to replace OD unit that has shown a poor performance in
urea rejection. Their results indicated that the new dual system did not only result in a complete urea rejection, but more
further, depends on the temperature gradient across the MD
membrane. The flux of DO/MD was 420 times greater than
in the DO/OD process with no sign of flux reduction due to
fouling.
The merits of MD capability to utilize low-grade waste
and/or alternative energy sources, such as solar and geothermal energy, made it more promising separation technique. For
instance, a possible area for MD implementation is the nuclear
industry or nuclear power stations, which are rich sources
in waste heat [22,29,103105]. Furthermore, solar-powered
MD technology is considered a feasible means for production
of pure water from brackish water in dry rural areas. Various solar pilot MD plants have been designed and proposed
[13,14,106109]. The tested solar-powered MD (SPMD) units
were found to be technically feasible. However, because the
variation of the daily solar radiation, the simulation results
showed that the most important factor in improving the plant
capacity is by adapting appropriate heat recovery systems
[109].

10

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

Table 2
Modeling of the MD process
Configuration

Areas of application

MD references

References in which modeling is


included

Type of model

DCMD

Best suited for applications in


which the major permeate
component is water and when
non-volatile components are
considered

[13,16,19,20,22,30,46,47,64,69
72,76,7880,82,83,85,89,9192,
100,106,109,111,112,117130,
136,139,140,154156,159,166]
(No. of references = 52)

AGMD

Because permeate fluids are not


in direct contact with the
membrane surface, the AGMD
can be applied to remove trace
volatile components from
aqueous solutions
For the removal volatile organic
components from aqueous
solutions
For the removal of volatile
organic components from
aqueous solutions

[13,8,9,16,19,20,22,29,30,33,43,
46,47,50,54,59,6165,6872,75
80,82,83,85,86,8994,97100,
101106,109,111,112,117131,
133136,138145,148157,159,166]
(No. of references = 96)
[1,2,5,7,13,14,24,25,3840,
66,73,74,107,108,110,114,116,
129,131,156,157,161,165] (No. of
references = 25)

[1,6,15,17,22,23,3436,41,42,
56,59,60,81,84,86,95,96,113,
131,158,160] (No. of references = 23)
[1,2,4,12,21,22,2628,115,
130132,162164] (No. of
references = 16)
160

[1,2,6,15,22,23,34,35,41,42,
56,60,81,84,113,158] (No. of
references = 16)
[1,2,4,21,22,2628,115,162164]
(No. of references = 12)

Most MD investigators
adapted the concept of
temperature and
concentration polarization.
The heat transfer models are
usually based on the
correlations originally derived
for heat transfer through
non-porous/rigid surfaces
(semi-empirical models). The
Mass transfer models based
on the dusty gas model, in
which the transmembrane
flux can take place via
contribution of Knudsen flow,
Poiseuille and molecular
diffusion flow. Combination
between these types of flows
haven also applied. In this
case the mean free path of the
transported molecules
through the membrane pores
in vapor phase is calculated
and compared to the
membrane pores size

VMD

SGMD

Total references

5. Modeling in MD
Various theoretical models have been developed for each MD
configuration [1]. The primary purpose of these models other
than the estimation of the temperature and concentration polarization coefficients is to predict the values of the permeate flux
and its dependence on the membrane module design, membrane
parameters and operating variables [16,15,16,2024,26,28,
30,3842,46,47,6974,78,8084,89,9192,110130]. Furthermore, other than the permeate flux, the membrane selectivity
was also predicted when using VMD configuration [84]. It is
worthwhile mentioning here that complete relationships on MD
modeling have been very well reviewed in Ref. [1]. Considering
other equations especially those involving the pore size distribution, which are not included in Ref. [1] are also well explained in
Refs. [80,81,83]. In MD both heat and mass transfer occur simultaneously and both temperature and concentration polarization
effects should be taken into consideration. Part of the transferred heat takes place by conduction through the membrane
matrix and the other part through the void space of the membrane
where gases are entrapped. Thus, the membrane parameters play
also an important role on the heat and mass transfer phenomena. Thus, all these factors result in complicated modeling steps.
Table 2 provides some of the related references found in literature that considered modeling in their MD studies. As can be
observed in Fig. 3 more than 61% of MD studies consider modeling as an essential part of their investigation. However, despite
the numerous mathematical models that have been developed by
MD investigators and regardless of their precision in predicting

[1,2,5,7,24,25,3840,73,74,
110,114,116,129,156,157] (No.
of references = 17)

97

the MD permeate flux, there are number of issues which are still
need to be fully understood.
It is worth mentioning that several similar mathematical models in MD have been presented. The analyses were based on the
assumption that gas permeates through a porous membrane comprising three contributions, namely, Knudsen flow, Poiseuille
flow, molecular diffusion flow and/or the transition between
them. The dusty gas model is frequently employed to calculate
the MD fluxes and the molecular diffusion flow is not considered
in VMD configuration. However the solutiondiffusion flow
through the non-porous portion of the membrane is not considered in spite of the existence of a high affinity (i.e. close
solubility parameters) between the species to be separated and
the membrane material, especially when VMD configuration
is considered [84]. In all MD models, a transport of adsorbed
molecules on membrane solid surfaces (i.e. surface diffusion) is
neglected because of the fact that the diffusion area of the membrane matrix is small compared to the membrane pore area [1].
For hydrophobic porous membranes with large pore sizes and
high porosity, as the affinity between water and the membrane
material is very low it may be allowed to neglect the contribution
of transport through the nonporous membrane portion. However,
the solution/diffusion mechanism may have a significant effect
once other compounds are present in the feed solution, especially
for species that have strong affinity with the membrane matrix
and for membranes having low porosity and small pore size. As
far as we know, within the membrane literature there is only
one paper dealing with this subject [84] using VMD configuration and pervaporation. Further investigation using other MD

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

configurations, membrane types and different volatile organic


compounds should be of great interests.
On the other hand, it must be mentioned that experimentally, in AGMD configuration, the permeate flux increases with
the increase of the membrane pore size. However, in this configuration, most of the developed theoretical models such as the
StefanMaxwell models do not include the pore size as a dependent parameter.
Numerical analysis using momentum, energy and diffusion equations have been made by various authors [4,110,112,
114,116,130]. However, these studies were not general in nature.
Most of the theoretical approaches consider an adjustment
parameter taken as the membrane pore tortuosity. In fact, the
value of this parameter must be determined experimentally.
One practical way is to evaluate the effective porosity, which
is the ratio of the membrane porosity and the pore tortuosity [17]. On the other hand, it is well known that the pore
size plays an important role when modeling MD as it permits
to elucidate the physical nature of the mass transport through
the membrane by comparison with the mean free path of the
transported volatile molecules. Nevertheless, the pore size distribution is considered only in very few mathematical models
[47,80,81,83,84,117,122,125].
It must be stated that the use of the empirical heat and mass
transfer correlations to predict the MD flux have been questioned. The theoretical curves presenting the MD permeate flux
as a function of the fluid circulation velocity were obtained
in some cases discontinuous due to changes from laminar to
transitional flow regime and from transitional to turbulent flow
[7,15,16,22,24]. It is worth quoting that the used heat transfer
empirical correlations were developed for non-porous and rigid
heat exchangers. Therefore, there is a difference between the
mechanism of heat transfer in MD systems, which is coupled
with mass transfer, and the mechanism of the heat transfer in
heat exchangers. Consequently, the applicability of the empirical heat transfer correlations in MD systems must be checked.
Another issue to be pointed out is the consideration of a
mean temperature and a mean concentration when modeling
MD process without taking into account the variation of these
parameters along the length of the membrane modules. In fact,
the hydrostatic pressure drop along the membrane module length
is needed to be considered and very few authors paid attention
to this point [75].
6. Operating variables affecting MD process
Table 3 shows the effects on the MD performance of the
different involved operating variables in each MD configuration
including the corresponding published papers.
6.1. Feed temperature
The effect of the feed temperature on permeate flux has been
widely investigated in the different cited MD configurations.
The feed temperature has been varied from 20 to 80 C (below
the boiling point of the feed solution) maintaining all other MD
parameters constant. Table 3 shows that generally it is agreed up

11

on that in all MD configurations there is an exponential increase


of the MD flux with the increase of the feed temperature. This is
due to the exponential increase of the vapor pressure of the feed
solution with temperature, which increases the transmembrane
vapor pressure (i.e. the driving force) as all the other involved
MD parameters are maintained invariables. It was stated that it is
better to work under high feed temperature as the internal evaporation efficiency, defined as the ratio of the heat that contributes
to evaporation and the total heat exchanged from the feed to
the permeate side is high although the temperature polarization
effect increases with the feed temperature [16,2123,47,129]. It
must be mentioned that when dealing with an aqueous solution
containing a volatile component, the exponential increase in permeate flux with the inlet feed temperature may be impeded by
the drop in selectivity. Such a phenomenon has been reported
by several MD investigators [6,23,26,27,40,41,132134]. This
maybe related to increased effect of both temperature and concentration polarization with the increase of feed temperature.
However through the tangential flow membrane modules, there
is a temperature and pressure drop. In most of the MD systems,
the inlet temperature is different from the outlet temperature,
which depends on the fluid circulation velocity. For this specified
reason it was suggested to adapt a local temperature polarization
coefficient value, and an overall one valid for the whole system
that may be defined and evaluated as a function of the experimental conditions [4,21]. One of the advantages of using stirring
cells (i.e. known as Lewis cells) is that the bulk temperature is
the same in the feed chamber as well as in the permeate chamber.
6.2. Feed inlet concentration
The effect of feed concentration on transmembrane flux
depends strongly on the separation process itself. As it is well
known MD can be used for the treatment of highly concentrated
solutions (i.e. non-volatile solute) without suffering the large
drop in permeability observed in other membrane processes such
as the pressure-driven membrane processes [1,5,75,91]. When
non-volatile solutes are considered, the most likely effect of the
feed concentration is to result in a decrease in the permeate flux
in all MD configurations (see Table 3). This is attributed to the
fact that the addition of non-volatile solute to water reduces the
partial vapor pressure and consequently reduces the driving force
of MD process. There is also the contribution due to the effect of
the concentration polarization (i.e. formation of a boundary layer
on the feed membrane surface). Nevertheless, this contribution
is very small compared to that of the temperature polarization
[6972,133]. On the other hand, when aqueous solutions containing volatile components (such as alcohols) are considered,
the effect of increasing solute concentration is completely different and depends upon the thermodynamic properties of the
involved volatile compound and its interaction with water. Generally, in such situations, increasing solute (volatile compound)
concentration results in a higher permeate flux. This is attributed
to the increase of the transmembrane partial pressure of the
volatile component due to the increase of its concentration in
the feed side. In this case, care must be taken to avoid membrane pore wetting. Another issue to be considered here is that

12

Table 3
Effect of operating variables on transmembrane flux in MD process
Configuration

Variable
Permeate side operating variables, effect of increasing of

Temperature

Concentration

Velocity

Stirring rate

Temperature

[13,1416,20,22,43,
47,61,64,6870,7375,
78,80,83,86,87,89,
91,99,104106,109,
117121,124,128130,
133135,149,151153,
156,157,159,166,167]

[1,22,47,61,64,69,70,
73,75,78,89,99,104,118,
119,121,129,131,134,135,
142,152,153,155,159]. But it

[1,15,16,43,61,69,70,78,86,
89,99,104,119,129,130,149,
153,154,156,159]

[3,83,89,118,
[47,80,89]
120,121,131,153]

[1,5,7,23,24,38,39,65,
107,114,116,131,156,
157,161]

[7,24,65,107]

Velocity

Stirring rate

[47,71,89,154]

[17,72,115,153]

Vapor pressure
difference

DCMD
[16,19,77,83,85,
98,99,134]

[22]

[22,47,118,131,152,155]
(for volatile solutes) [133,166]

AGMD
[5,7,23,24,38,39,65,107,
114,116,131,156,165]

[5,114] or

[5,23,39]

[107,114]

a
[5,23,39,114] (cooling
water velocity)

[5,23,38,73]

[4,115]

[7,38,39]
SGMD
[1,4,21,22,26,27,28,
115,132,162,164]

[28]

[132,164] or

[4,21,115] or

(for volatile solutes) [132,164]

[4,21,22,28,115]

[132,162]

[6,94,96,160]

[6,15,23,42,94,130,158,160]
[6,15,23,42,94,130,158,160] or

[4,21,22,26,27,28,
115,132,162,164]
(sweeping gas velocity)

VMD
[1,15,17,23,41,
84,86,9496,113,
130,149,158, 160]

[17,23,41,84][17,23,41,84]
(for volatile solutes)

[17][17]

[6,17,41,42,113]
[6,17,41,42,113]

[22,41,84,86,95,96,149]
[22,41,84,86,95,96,149]

: increases with; : decreases with; : increases but reach asymptotic level; : there is an optimum value;
a With regards to the cooling water.
b Related to the sweeping gas.

: not clear or not relevant;

: slight increase or decrease with. Bold numbers for volatile solutes.

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

Feed side operating variables, effect of increasing of

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

the presence of a non-volatile component such as salts in diluted


aqueous solutions containing organic compounds may alter the
vapor pressure of the solution. For instance in case of a binary
mixture of water/alcohol, the addition of salt may change the
vapor pressure ratio and increases the MD selectivity [40,132].
6.3. Feed circulation velocity and stirring rate
The effect of both the feed circulation velocity and feed stirring rate is to increase the heat transfer coefficient in the feed
side of the membrane module and to reduce the temperature
and concentration polarization effects, thus resulting in higher
MD permeate fluxes. It was observed in some MD systems
that the MD flux increases with the feed circulation velocity
and stirring rate and tends to an asymptotic value at higher
feed flow rates [2,5,22,39,40,66,116,131]. However, other MD
studies show only a practically linear increase of the MD flux
with the feed circulation velocity [15,16,22,86,132]. To obtain
a higher productivity it is better to operate under turbulent
flow regime [23,25,50,70,78,86,91], which is accomplished by
employing higher mixing intensity or higher circulation velocity. This means that the temperature at the membrane surface
become closer to the bulk feed temperature and therefore the
transmembrane temperature difference is higher. Table 3 shows
this general effect of feed flow velocity for each MD configuration. However, it can be seen that the response may differ
from one to another MD configuration and MD set-up. While
the DCMD, AGMD and VMD configurations show an increase
of the permeate flux with the feed flow rate, reaching in some
cases asymptotic levels, the effect of the feed flow rate on the
SGMD flux is practically negligible when non-volatile solutes
were present in the feed solutions [4,22,115]. This means that
the temperature polarization in the feed side of the membrane
module in SGMD has a little effect as it is localized in the permeate side [4,21,22,28,115]. It is worth pointing out that when
volatile components are present in the feed solutions, the process separation factor (selectivity) is most likely to be directly
affected by the feed solution flow velocity. Higher separation
factors were reported by various MD investigators in almost all
MD configurations [6,39,41,42,132,164]. The enhancement of
the process separation factor is related to the better mixing condition at higher feed flow velocities, which decrease the effect
of the concentration polarization boundary layer. It is worthwhile mentioning that there is a linear relationship between the
Reynolds number (ReN ) and the fluid velocity (u), whereas the
pressure drop (P) along the length of the membrane module
(L) is a function of the second power of the fluid velocity.
ReN =

udh

P = f

L u2

dh 2

(2)
(3)

where dh is the hydraulic diameter of the flow channel, f friction factor, and are the fluid density and dynamic viscosity,
respectively. Therefore, the feed flow rate must be varied with
due precautions to avoid membrane pore wetting as the trans-

13

membrane hydrostatic pressure must be lower than the liquid


entry pressure of feed solution (LEP) into the membrane pores
and to assure working under turbulent flow regime.
It must be mentioned that although stirring cells (i.e. Lewis
cells) are not suitable for MD industrial implementation, these
type of MD set-ups demonstrate to be useful for laboratory
DCMD experiments [3,89,118,121]. In this type of MD systems, enhancement of the DCMD flux with stirring rate up to an
asymptotic value is easily observed. Table 3 shows that no study
was found in which stirring cells were used either in AGMD or
in SGMD.
6.4. Permeate inlet temperature
The general effect of increasing the permeate temperature is
more likely to result in lower permeate flux. This is related to the
decrease of the transmembrane vapor pressure as far as the feed
temperature is kept constant. Table 3 shows that the effect of
decreasing the permeate temperature is to increase the permeate
flux for DCMD applications [1,2,22,47,91,125]; however there
is a negligible effect of permeate temperature (temperature of
the cooling plate) rise in the case of AGMD [5,24]. In this configuration, the heat transfer coefficient in the air gap dominates
the overall heat transfer coefficient and therefore changes of the
permeate temperature will have little effects on the AGMD flux
[5,24,40]. The situation is quite similar in the case of SGMD
as the temperature of the sweeping gas in the permeate side
increases very fast from the inlet to the outlet of the membrane
module and the effect on the permeate flux is negligible. As a
consequence, it is more convenient to increase the feed side temperature rather than to decrease the permeate side temperature
in order to obtain higher MD flux.
As stated earlier, the permeate temperature increases considerably along the membrane module length especially in the
case of SGMD configuration in which the difference between air
inlet and outlet temperatures approached values as high as 20 C
[26]. This variation depends strongly on the fluid type and on its
flow velocity affecting the driving force and the transmembrane
flux [4]. In order to overcome this problem in SGMD, a cold
wall is introduced at the permeate side to retain the sweeping
gas temperature and to maintain the driving force as constant as
possible along the membrane module [26,27].
6.5. Temperature difference and mean temperature effect
The driving force in MD is the transmembrane vapor pressure, which can be applied in the case of DCMD, AGMD and
SGMD by a temperature difference between the feed and permeate side of the membrane module. When the mean temperature
was maintained constant, it was observed that the permeate flux
increases linearly with the temperature difference [13,43,161].
On the other hand, when the temperature difference was maintained fixed, the permeate flux was found to increase exponentially with the mean temperature [2,19,21]. An Arrhenious type
of dependence is frequently considered to fit the obtained experimental data [2]. Similar results have been reported by other
MD investigators in AGMD modules [2,25,39,116]. However,

14

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

a different trend was observed from experiments conducted on


SGMD modules [4,27]. The transmembrane flux was found to
increase monotonically with the increase in bulk temperature
difference. This conversely trend may be related to the unavoidable temperature and concentration polarization effects. It is
worthy to be mentioned, as observed in Ref. [134], that the tendency of the transmembrane flux increase with the increase of the
bulk temperature difference depends upon which stream temperature is kept constant. If the permeate stream temperature is kept
constant, while increasing the feed side temperature, a monotonic increase in flux with respect to the increase in bulk phase
temperature difference would be observed. On the other hand, if
the feed stream temperature is kept constant, while varying the
permeate stream temperature, it is more likely to have an increase
in transmembrane flux tending to approach an asymptotic value
at higher bulk phase temperature differences [47,91,134].
6.6. Permeate ow velocity
The effect of the permeate flow velocity can be investigated
only for DCMD and SGMD configurations. An increase of the
permeate flow velocity increases the heat transfer in the permeate side of the membrane module by reducing the temperature
and concentration polarization effect if volatile components are
present in the permeate side. As the heat transfer coefficient in
the permeate side increases, the temperature at the membrane
surface approaches the temperature in the bulk permeate side
and consequently the driving force as well as the permeate flux
increases as can be seen in Table 3. However, as the permeate
fluid type in the DCMD is more likely distilled water and in the
SGMD is usually humid air, the extent of the velocity effect is
much more pronounced in the later configuration. This may be
related to the different physico-chemical properties of these two
fluids and precisely to the thermal conductivity between liquids
(water) and gases (air). In the case of SGMD, when non-volatile
components are present in the feed solution, it was confirmed
that the temperature polarization is localized at the permeate side
[4,21,22,28,69,115,164] and there is an optimum permeate flow
velocity beyond which any increase in permeate velocity may
result in a negative effect or flux decline. Therefore, the gas flow
in SGMD should be optimized to achieve the maximum water
transmembrane flux. This result was attributed to the increase of
the permeate pressure. In fact, the gas flow must be varied with
due precautions, the transmembrane hydrostatic pressure must
be lower than the liquid entry pressure of the feed solution (LEP)
and the permeate pressure must be lower than the pressure in the
feed side.

tions. Due to the fact that the temperature and the concentration
polarization affect the permeate flux, it was also observed a
non-linear relationship between the permeate flux and the bulk
pressure difference [2]. In VMD both the permeate flux and the
transmembrane hydrostatic pressure increase with the decrease
of the vacuum pressure applied in the permeate side (downstream) and therefore the risk of membrane pore wetting is very
high. Nonetheless, when a volatile organic compound (VOC) is
present in the aqueous feed solution, the effect of lowering the
downstream pressure in VMD may result in higher total permeate flux (water and VOC) but in contradiction it may also
lead to a relatively poor separation factor of the process. To
achieve highly concentrated permeate with VOC, it is recommended to work at permeate pressures (downstream pressures)
higher than the vapor pressure of water [35,41,42]. On the other
hand, the effect of the transmembrane hydrostatic pressure in
MD configurations is not clear and needed to be studied in all
MD configurations.
7. Membrane parameters affecting MD process
The membranes used in MD process are actually made
for microfiltration purposes and most of them are prepared
using polypropylene (PP), polyvinylidene fluoride (PVDF)
and polytetrafluoroethylene (PTFE). These membranes are
available in capillary or flat-sheet forms [128]. The required
membrane parameters for MD process were given in details
by many authors [1,9,17,18,93,104,135]. In comparison to
other membrane separation processes, only few authors have
considered the possibility of designing membranes for MD
[1719,8691,9395]. A good porous membrane should exhibit
low membrane resistance to mass transfer, high liquid entry pressure of water to maintain the membrane pores dry, low thermal
conductivity to prevent heat loss through the membrane matrix,
good thermal stability and excellent chemical resistance to most
of the feed solutions. The relationship between the transmembrane flux and the different membrane characteristic parameters
is given in Ref. [1] as
N

ra
m

(4)

where N is the molar flux, r the mean pore size of the membrane
pores, a factor whose value equals 1 or 2 for Knudsen diffusion
and viscous fluxes, respectively,m the membrane thickness,
the membrane porosity and is the membrane tortuosity. Table 4
presents the effect of each membrane parameter on the MD permeate flux for each MD configuration.

6.7. Vapor pressure difference

7.1. Membrane thickness

As stated earlier, the MD driving force is the vapor pressure


difference between both membrane sides. This can be applied by
a temperature difference or by applying vacuum in the permeate
side of the membrane module. Table 3 shows the obvious effect
of the transmembrane vapor pressure on the permeate flux. It is
generally admitted the linear increase of the MD flux with the
transmembrane vapor pressure difference in all MD configura-

As occurred in other membrane separation processes, the


permeate flux in MD is inversely proportional to the membrane thickness. Membrane thickness plays a significant role
in dictating the resistance to mass transfer. To obtain a high
MD permeability, the membrane should be as thin as possible.
On the contrary, to achieve better heat efficiency the membrane
should be as thick as possible due to the fact that in MD heat

: not clear or no relevant reference was found. Bold numbers


: slight increase/decrease with; : there is an optimum value;
: increases with; : decreases with; : very important may increase or decrease;
for volatile solutes.

[81,160]
[81,84][81,84]
[17,23,42,81,84,95,115]
[17,23,42,81,84,95,115]
[17,23,42,81,84,95,115]
[17,23,42,81,84,95,115]
[7,81,86]
VMD

SGMD

[1]

[39,157][39,157]

[4]

[7,39,114][7,39,114]

[4,21,115]

[5,7,25,39][5,7,25,39]

[80]
[80,82,157]
AGMD

[1,47,61,78,89,135]
DCMD

[95,14]

[5]
[39][39]

[81,115]

[43,47,75,91,118,
121,135,153]
[1,18,61,62,75,89,90,
94,167][1,18,61,62,75,
89,90,94,167]
[47,78,91,117]
[47,82,117]
[1,7,75,78,86,89,91,
117,128,151]
[1,86,91,117,135]

Pore size
distribution
Pore size
Porosity
Membrane
thickness

Parameter
Configuration

Table 4
Membrane parameters affecting permeate flux in MD process

Surface geometry
Surface chemistry
Tortuosity

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

15

loss by conduction takes place through the membrane matrix


[1,3,22,47,78]. By the aid of a computer simulation programme,
considering the thermal conductivity of the commercial membranes, the optimum thickness has been estimated and found to
be within the range of 3060 m [47]. However, in DCMD it was
confirmed the use of porous composite hydrophobic/hydrophilic
membranes with very thin hydrophobic layer, as low as 5 m
[89,136].
The collected results summarized in Table 4 show that as
the membrane thickness is higher, the permeate flux gets lower.
This is a general trend in all MD configurations except in AGMD
configuration, which shows that the influence of using thicker
membranes seems negligible due to the predominant resistance
to mass transfer provided by the stagnant air gap [5,24,25,40].
This fact must be carefully studied by using membranes with different thicknesses but similar pore size and porosity. In addition,
MD studies have appeared investigating the effect of membrane compaction although MD is not a pressure-driven process
[47,124,137].
7.2. Membrane porosity
Membrane porosity also termed membrane void volume is an
important parameter affecting MD flux. Membranes with higher
porosity exhibit greater surface area for evaporation. In general,
the MD membrane porosity lies between 30 and 85%. As can
be seen in Table 4, it is generally agreed upon that higher membrane porosity, results in higher permeate fluxes regardless of
which MD configuration is used. It must be pointed out that
those membranes with high porosity exhibit lower conductive
heat loss since the conductive heat transfer coefficient of the
gases entrapped within the membrane pores is an order of magnitude smaller than that of the hydrophobic polymer used for
membrane preparation [1,3,72,79,89].
7.3. Membrane pore size
The membranes used in MD exhibit pores sizes ranging from
100 nm to 1 m and it is admitted that the MD flux increases with
the increase of the pore size. The increase in transmembrane
flux with increasing membrane pore size may be related to the
enhanced mass transport process from being more likely Knudsen diffusion controlled for membranes having very small pore
sizes to Knudsen-viscous transition for membranes exhibiting
larger pore size [75,135]. A rigorous systematic study on this
subject is not investigated yet. In order to avoid wettability of
membrane pores, the pore size must be as small as possible
leading to a conflict with the requirement of higher MD permeability. An optimum value is needed to be determined for each
MD application depending on the type of the feed solution to
be treated. Based on the Kinetic Theory of Gases, by comparing
the membrane pore size to the mean free path of the transported
volatile molecules through the membrane pores, the mechanism
of mass transfer responsible of MD flux can be defined and
its magnitude predicted and compared with the experimental
data.

16

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

7.4. Pore size distribution


The membranes employed in MD exhibit a pore size distribution rather than a uniform pore size. Therefore, more than
one mechanism can take place simultaneously to different extent
depending on the pore size and on the MD operating conditions.
The effect of pore size distribution on the MD flux has been rarely
studied in DCMD configuration [47,80,83,113,117,122] and in
VMD [81,84]. As can be seen in Table 4, the effect of membrane
pore size distribution is not studied yet in SGMD and AGMD.
It was observed when using commercial membranes that the
predicted DCMD flux assuming uniform pore size is quite similar to that calculated using pore size distribution. In contrast,
when using laboratory made membranes, which exhibit pore size
distribution with geometric standard deviations far from unity,
the difference between the calculated permeate fluxes is higher
[117]. Further studies are needed to clarify this effect using other
membranes and MD configurations.
7.5. Pore tortuosity
Membrane tortuosity is the average length of the pores compared to the membrane thickness. In MD studies and in order
to predict the transmembrane flux, a value of 2 is frequently
assumed for tortuosity factor [1,3,41,80,81,138]. Furthermore,
a tortuosity factor value as high as 3.9 was reported [124]. Generally, the membrane pores do not go straight across the membrane
and the diffusing molecules must move along tortuous paths,
which lead to lower MD fluxes. In fact, the MD flux is inversely
proportional to the membrane thickness times its tortuosity. To
our knowledge, no systematic study has been performed on the
effect of pore tortuosity on the MD flux (Table 4). Due to the
difficulties in measuring its real value for any of the microporous membranes used in MD process, investigators frequently
assume it as a correction factor helping on the prediction of the
MD flux and therefore hiding other aspects that are worth to be
investigated further. Nonetheless, the gas permeation test combined with the measurement of the porosity permits to estimate
this parameter [81,117,125]. It must be mentioned that membranes prepared by means of ionic bombardment exhibit very
low pore tortuosity [104].
7.6. Membrane surface chemistry
Generally, the membrane in MD acts only as a support of
the vapor/liquid interfaces and does not modify the vapor/liquid
equilibrium of the feed aqueous solutions. However, the membrane must be resistant to the feed solutions. One of the major
requirements of MD process relies on the use of hydrophobic
membranes to maintain dry the membrane pores. Hydrophobicity can be achieved by either using hydrophobic materials such
as polypropylene (PP), polyethylene (PE), polytetrafluoroethylene (PTFE) and polyvinylidene fluoride (PVDF) or by making
the membrane surface energy as low as possible by means of
different surface modification techniques of hydrophilic membranes [18,89,90,136,168]. In a recent development in membrane science, the surfaces of hydrophobic microporous mem-

branes (facing hot solution) were protected against wet-out by


a new modification technique [94,95,149]. The new and novel
technology involves a highly water permeable plasma polymerization coating of silicon fluoropolymer layer on the outside
surface of porous hydrophobic membranes (PP fibers). The
extremely thin layer is porous itself and covers the entire surface
(brine side) to create a patchy layer of hydrophobic microporous/porous coating on the surface of hydrophobic porous
membranes [94,95,149]. It is worth quoting here that the surface of hydrophobic porous membranes were also modified
by a hydrophilic polymer generating thin hydrophilic porous
layer in order to protect the membrane pores against wetting by
surface active agents [167]. The hydrophobic microporous flat
sheet membrane made of PVDF (used as substrate here) was
first treated in a potassium dichromic-sulphuric acid solution,
washed with plenty of water and left to dry. A prepared polymer
gel was then cast on the surface of the hydrophobic substrate
to form the composite membrane. The composite membrane
with a PVA/PEG hydrophilic layer claimed to prohibit the wetting problem exhibited by traditional hydrophobic membranes
employed in MD [167]. As can be seen in Table 4, compared
to the number of reported studies on MD, very few works have
been carried out on the design of membrane for MD. The effect
of the membrane material on the selectivity of MD process
is needed to be deeply investigated especially when volatile
organic compounds are present in the feed solution and when
thick membranes with low porosity are used. In other membrane processes, membrane roughness affects the concentration
polarization, the membrane fouling effect and the permeate flux
together with the membrane selectivity. This membrane roughness effect on the MD flux was mentioned recently and is needed
to be deeply studied in all MD configurations [81].
7.7. Membrane module geometry
Flat-sheet membranes in plate and frame modules or spiral
wound modules and capillary membranes in tubular modules
have been used in various MD studies. The design of MD modules must permit high feed and permeate flow rates with high
turbulence and low pressure drop along the membrane module. The availability of the industrial MD modules is up to now
one of the limitations for MD process implementation. Due to
the fact that MD is a non-isothermal process, the design of the
MD modules must not only provide good flow conditions, low
pressure drop and high packing density but must also guarantee a good heat recovery function and thermal stability. A spiral
wound module with integrated heat recovery has been proposed
for AGMD integrated with solar-powered desalination system
[13,14]. At laboratory scale, high flux using improved DCMD
modules have been reported [73,75,91,92,94,95,149]. Another
important aspect is the deficiency in fluid flow along side the
effective mass transfer area, that is to say the flow maldistribution
mainly affecting hollow fiber module in MD. Flow maldistribution is caused by the polydispresity of fiber inner diameter at the
lumen side, while at the shell side it is due to the non-uniformity
of fiber packing [126]. It has been reported that the effect of
the randomness of fiber packing in shell side can lead to trans-

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

membrane flux reduction by up to 58% (in comparison to ideal


module with uniform packing) when a 0.4 packing fraction was
used. However, increasing the packing fraction can minimize
its effect [126,139,140]. One major advantage of hollow fiber
modules over other MD modules is its high rate of mass transfer
which is a result of its high surface density area.
8. Framework for better understanding MD process
The important influence of the different operating variables
and membrane factors on the transmembrane MD flux and the
enhancements achieved so far in this field have been reported.
The question yet to be answered is: Why the MD process has
not yet been industrially implemented on a large scale?
In addition to the earlier barriers mentioned previously, some
of MD authors have referred this to the fact that MD technology has come at a time when industries are cutting back
and are not willing to invest in emerging processes, which
may be competitive to other used techniques [1]. Some others have related this to insufficient understanding of temperature/concentration polarization and surface membrane fouling
[73]. Other MD investigators claimed that the understanding in
heat transfer of MD is still incomplete [16] and others attributed
this to the design and development of membranes for MD
[1719,86,89,90,9395,149] and modules [7377,95,135,149].
However, before MD can be industrially implemented various
MD issues have to be deeply investigated including the longterm MD performance, MD membrane fouling, MD membranes
and modules engineering, energetic and maintenance cost analysis, some aspects in theoretical heat and mass transfer models
especially there is a need of a systematic study on the effect of
membrane pore size on the change of the MD type of flow as
well as the effect of the pore tortuosity on the type of flow in all
MD configurations.
Fig. 4 illustrates the sequences of the different steps that
should be followed to fulfill the requirements for any MD industrial application. Attempts are made in Fig. 4 to address the
different sequences and steps required towards possible industrialization of the MD process and Table 5, on the other hand,
to highlight the different aspects of the process showing those
areas that are either rarely studied or far from being understood.

17

MD configuration. Each of the known MD configurations may


suit different application areas as stated in Tables 1 and 2. The
MD process can be performed with any of the module types that
have been mentioned in Section 7.7. However, one has to choose
the most suitable for the proposed application. Selection of the
appropriate membrane with the required characteristics is very
important, as this would directly affect MD process. The important influence of each parameter characterizing the membrane
has been discussed earlier in Section 7 of this study. The effect
of different operating conditions on the MD process have been
earlier discussed and summarized in Table 3.
It is worth pointing out that knowledge about physical and
chemical properties of the feed solution to be used is necessary in order to select the right MD operating conditions.
Nevertheless, depending on the obtained experimental results,
the operating conditions may be changed accordingly. In order
to get a complete sight of the applicability of MD process to
the aimed separation, one should investigate the possibility of
fouling occurrence and MD long-term performance. Whether
fouling or wetting has been detected, the possibility of diminishing or minimizing their effects should be investigated. If it is
succeeded then a feasibility study may be launched to check if
the process is economically viable.
Prior to the design of an industrial MD unit one must be
sure of its viability through scaling up and the design of a pilot
plant unit. The pilot plant scale unit should be tested taking into
consideration the optimized operating conditions that have been
achieved during the laboratory experiments. If similar results
and performance are obtained then it would be save to design
an industrial scale MD production unit. Otherwise, depending
on the deficiency encountered, changes should be made in any
of the operating conditions and variables to achieve the required
results. One may not rule out the possibility of coupling MD
process with other existing technologies as this has been found
useful by many MD investigators [13,14,65,97102,106109].
8.2. The different aspects of the MD process
In order reach the industrial implementation of the MD process, all aspects of MD should be fully understood. Table 5 is an
attempt to highlight areas in MD that are either usually or rarely
studied and the degree of their understanding.

8.1. Flow chart towards the industrialization of MD process


The increasing growing interest in MD has lead to a wide
variety of areas that suit its application. However, despite of all
achieved encouraging results none of them is industrially implemented. Fig. 4 presents a flow chart showing the different steps
that may be followed towards possible industrialization of the
MD process in different areas. First it must be known that MD
can be applied generally for all aqueous solutions having different non-volatile solutes under a wide range of concentration
with a separation factor near 99.9%, and for dilute organic aqueous solutions in a reduced range leading to variable separation
factors depending on the feed solution to be treated, membrane
and operational conditions. Having chosen the MD process for
a given application, the first step is to decide the most suitable

8.2.1. Heat transfer


The mechanism of transport in MD involves simultaneously
both heat and mass transfer. The aspect of heat transfer in all
MD configurations is very important and is more likely believed
to be the rate controlling steps in the MD process. It is carried
out in four steps: (i) heat transferred from the feed solution to
the membrane surface across the thermally boundary layer in
the feed side of the membrane module; this is related to the
temperature polarization effect; (ii) heat transport by conduction across both the membrane matrix and the gas filled pores
considered heat loss in MD; (iii) heat associated to the latent
heat of vaporization and therefore to the mass transfer through
the membrane pores (the efficient heat in MD); (iv) heat transfer
from the membrane surface to the permeate solution across the

18

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

Fig. 4. Flow chart towards industrialization of MD process for any proposed application.

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

19

Fig. 4. (Continued ).

thermally boundary layer in the permeate side also associated to


the temperature polarization effect.
8.2.1.1. Heat transfer through feed side boundary layers (temperature polarization effect). As in any other process involving
heat transfer, heat flows from the higher to the lower temperature region. The more the temperature difference is the more
the rate of heat flow would be. The heat transfer boundary layer
formed at each side of the membrane surface imposes a resistance to heat transfer and makes the temperature difference at the
liquid/membrane interfaces lower than that applied at the bulk
phases. This affects negatively the driving force for mass transfer. It is generally agreed upon among MD investigators that one
real reason for low fluxes in MD process is due to the effect of
temperature polarization. In fact, up to 80% reduction in driving
force may be observed due to the temperature polarization effect
[1,6971,74]. In fact, the term temperature polarization coefficient (TPC) was defined as the fraction of the transmembrane
temperature to the bulk temperature difference.
TPC =

Tfm Tpm
Tfb Tpb

(5)

where Tfm , Tpm , Tfb and Tpb are membrane surface temperatures and fluid bulk temperatures at the feed and permeate side,
respectively. A schematic diagram of the temperature polarization in MD is shown in Fig. 5. It can be seen that due to the
temperature polarization effects, the bulk feed temperature Tfb
is gradually decreased across the developed thermal feed boundary layer of thickness ft to Tfm , which is the feed temperature at
the feed/membrane surface. Similarly, at the permeate side (not
applied to VMD configuration), the temperature at membrane
surface (Tpm ) is higher than that of the permeate bulk phase
(Tpb ) due to the developed thermal boundary layer of thickness
pt . It is worth mentioning that both feed and permeate thermal
boundary layers are function of fluid properties and operating
conditions, as well as the hydrodynamic conditions. It is convenient to work under optimum mixing conditions to highly
diminish the temperature polarization effects. Enhanced heat
and mass transfer can be achieved by improving the design of
flow passage, membrane arrangement or by applying turbulence
promoters like mesh spacers. Working at high flow velocities is
always an option that is usually considered by MD investigators
to achieve better mixing conditions and minimizing the effect
of temperature polarization [23,75,91,95]. However, it will be

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

Particulate and
corrosion fouling

Energy consumption
and production cost
evaluation

Fouling type

Crystallization
fouling

Biological
fouling

20

VMD

SGMD

AGMD

DCMD

: most studied areas; : fairly studied areas; : least studied;

Conduction Convection

Mass transfer across


boundary layers
(concentration
polarization)
Heat transfer across the membrane

Heat transfer across


boundary layers
(temperature
polarization)

Heat transfer

Configuration Aspect

Table 5
Inadequate understanding of the MD phenomenon

Mass transfer

Mass transfer
across the
membrane

: scattered studies or unstudied area.

Module design
(surface geometry
and operation
mode)

Process rate
limiting step

Surface chemistry and


membranes for MD
processes

Long-term
performance

Fig. 5. Temperature and concentration polarization in MD.

economically feasible to work under optimum conditions. In


this case, it is necessary to ensure low pressure drop along the
membrane module in order to avoid membrane pore wetting
especially when membranes of large maximum pore size and
organic aqueous solutions are employed. It has been shown earlier in Eq. (3) that the pressure drop along the length of the
membrane module is a function of the second power of the fluid
velocity. Thus flow velocities should be always increased with
precautions.
It must be stated here that the TPC may be related to the heat
transfer coefficients involved in the MD system [1 add other
references also]. Since MD relies on phase change to perform
the required separation process, at least the latent heat for evaporation must be continuously transferred from the bulk feed to
the membrane surface. The supplied heat flux, qf , depends on
the film heat transfer coefficient of the boundary layer, hf , and
the temperature difference between the feed bulk and membrane
surface as follows if the heat lost to the environment is not considered.
qf = hf (Tfb Tfm )

(6)

It was found that the temperature polarization effect increases


with the feed temperature although higher feed temperature usually results in an exponential increase in the transmembrane flux
[16,21,22,47,71,129].
8.2.1.2. Heat transport by conduction through the membrane.
In MD, the fraction of heat transferred by conduction through
both the membrane matrix and the gas filled pores is considered
heat lost and should be minimized in order to decrease the temperature polarization effect and increase the efficiency of the MD
process. Unlike other MD configurations, the internal heat lost
by conduction in VMD is usually considered negligible due to
the existence of low vacuum pressure at the permeate side. The
following expression may be used for the heat flux transferred
by conduction through the membrane.
qm = hm (Tfm Tpm )

(7)

where hm is the heat transfer coefficient, which can be evaluated


as follows:
m
hm =
(8)
m

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

where m is the average thermal conductivity of the membrane,


which takes into account both thermal conductivity of the membrane matrix and thermal conductivity of the gas present in the
membrane pores:
m = s (1 )g

(9)

where is the membrane porosity, s and g are the thermal


conductivities of the membrane material and the gas in the pores,
respectively.
It must be pointed out that the thermal conductivity of
air/gases is an order of magnitude lower than that of the membrane material. Therefore, the heat lost by conduction through
the membrane can be minimized by using membranes with high
porosity (i.e. high void volume).
8.2.1.3. Efcient heat in MD. Since MD relies on phase change
to perform the required separation process, the amount of heat
that contributes in the evaporation step is considered the efficient
heat. The efficiency of the MD process will be maximized if the
temperature polarization effect and the internal heat lost by conduction through the membrane and the external heat lost to the
environment are reduced. Depending on the way that the latent
heat of vaporization is being supplied, mass transport of the
evaporated species takes place from the vapor/liquid interface
formed at the feed/membrane pore interface. The corresponding
heat flux can be estimated by the following expression.
qv = NHv

(10)

where N is the MD molar flux and Hv is the latent heat of


vaporization.
The total heat transferred through the membrane whether it
is considered efficient heat or heat lost by conduction may be
written as follows:
qm+v = qm + qv = hm (Tfm Tpm ) + NHv

(11)

8.2.1.4. Heat transfer in permeate side. The heat transfer from


the membrane surface to the bulk permeate side, across the
thermally boundary layer, is also associated to the temperature
polarization effect. Fig. 5 shows that due to the temperature
polarization effect, the temperature at the membrane surface of
the permeate side is higher than that at bulk permeate. This is
more likely the case in all MD configurations except VMD due
to the applied vacuum in this side of the membrane. The temperature polarization effects in the permeate membrane side may
considerably diminished in similar ways applied to the feed side.
However, in AGMD configuration in which a stagnant air/gas
gap exists between the membrane and the cooling surfaces, the
temperature polarization effect at the permeate side can only
be reduced by optimizing the air/gas gap width. The heat flux
that takes place across the thermal boundary layer thickness at
permeate side may be given by the following equation:
qp = hp (Tpm Tpb )

(12)

where hp is the heat transfer coefficient of the thermal boundary layer at permeate side. Under steady state conditions, the

21

following equation is considered:


qf = qm+v = qp

(13)

Generally speaking, in MD the heat transfer coefficients of


the boundary layers are usually estimated from the well-known
heat transfer empirical correlations that relate the dimensionless Nusselt number with the Reynolds and Prandtl numbers.
Those correlations have been obtained for non-porous rigid heat
exchangers. On the contrary, the membrane surface is porous and
not rigid involving mass transfer. Thus, some MD investigators
have questioned their applicability in modeling MD processes
[15,16]. Moreover, Table 5 shows that one of the major studied aspects in MD process is heat transfer. However, mainly
the DCMD configuration is the subject of investigation by MD
researchers. The heat transfer aspect in MD is far from being
fully understood especially in the SGMD and VMD configurations and much works in these fields are needed to check the
applicability of the heat transfer empirical correlations.
8.2.2. Mass transfer
Fig. 5 illustrates also different situations of mass transport
depending on the volatility of the solutes considered in the aqueous feed solutions. The mass transport of the volatile species is
taking place from (i) the bulk feed to the membrane surface; (ii)
transported through the membrane pores in gaseous phase; (iii)
and then from the membrane surface at the permeate side to the
bulk permeate phase.
8.2.2.1. Mass transport trough feed side boundary layers (concentration polarization). When a single volatile component is
considered in the feed side, there would be no concentration
gradient (CAb = CAm ) only temperature polarization effect is
applied. On the contrary, as shown in Fig. 5 when more than
one component is present in the feed aqueous solution both concentration and temperature gradient occur simultaneously. In
the first case, when aqueous solutions containing non-volatile
solute(s) are considered, the concentration of the non-volatile
solute(s) at the membrane surface (CBm ) becomes higher than
that at the bulk feed (CBb ) with time as long as the separation
process is taking place. Almost 100% of separation is obtained.
In this case, care must be taken as supersaturation states may
eventually be achieved affecting the efficiency of the membrane
process.
The increased concentration of non-volatile compound next
to the membrane surface would have the influence of reducing
the transmembrane flux due to the establishment of concentration polarization (CP) layer of thickness, fC , at the feed side that
acts as a mass transfer resistance to the volatile molecule species
(water). In other membrane separation process (pressure-driven)
such UF, concentration polarization is usually considered a
major cause for flux decline [1,3,7072]. Fortunately, in MD
process, the low to moderate flow rates and high heat transfer
coefficients reduce its impact, which is lower than that of the
temperature polarization effect [70,72]. The term concentration
polarization coefficient, CPC, is defined to quantify the mass
transport resistance within the concentration boundary layer at

22

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

the feed side as follows:


CPC =

CBm
CBb

vapor pressure difference across the membrane


(14)

The molar flux of species A (volatile component) through the


concentration boundary layer of non-volatile component B may
be calculated from the following mass balance equation [1]:


CBb
NA = CkAf ln
(15)
CBm
where NA is the molar flux of volatile component A, C the bulk
liquid phase total molar concentration (CA + CB ) and kAf is the
mass transfer coefficient of the volatile component through the
concentration boundary layer.
The second case shown in Fig. 5 is for the separation or
removal of volatile compounds from aqueous solutions. In this
case, the aim is either to have a permeate side of higher solute
concentrations (volatile component) or feed side free from any
traces of volatile components. Unlike the previous situation in
which the non-volatile component (solute) imposes the resistance to mass transport (CP boundary layer), separation of aqueous solutions containing volatile components usually is associated with concentration boundary layers made of the solvent (A,
water) and not the solute. Fig. 5 shows that the concentration of
component R decreased across the concentration boundary layer
of thickness fC from CRb at the bulk feed phase to CRm at the
membrane surface. The molar flux of volatile component, NR
across the concentration boundary layer (solvent, usually water,
considering a binary mixture) can be expressed as


CAb
NR = CkRf ln
(16)
CAm
where C is the bulk liquid phase total molar concentration
(CA + CR ) and kRf is the mass transfer coefficient of volatile
component R across the concentration feed boundary layer. It
should be emphasized that when dealing with aqueous solutions of volatile components, the separation process is usually
associated with transport of both the volatile component and
the water molecules into the permeate side, thus altering the
process separation efficiency. The best way to achieve a high
efficient separation process is by combination of selecting the
proper membrane characteristics and good operating conditions
that would favor the transport of volatile component species over
that water.
8.2.2.2. Mass transport through the membrane pores. As stated
earlier, the mass transfer process in MD is driven by the imposed
vapor pressure gradient between both sides of the membrane.
The mass transport mechanism is governed by three basic mechanisms known as Knudsen-diffusion (K), Poiseuille-flow (P),
Molecular-diffusion (M) or the combinations between them
known as transition mechanism. The Dustygas model is usually
used as a general model taking into account the latter basic mechanisms [1,3,22,80,81,83,84,91,115,137]. Regardless of which
mechanism is involved in the mass transportation process, the
models suggest that the molar flux, N, is proportional to the

N = C[Pfm (Tfm ) Ppm (Tpm )]

(17)

where C is the membrane distillation coefficient that can be


obtained experimentally, Pfm (Tfm ) and Ppm (Tpm ) are the vapor
pressure as function of temperature at both the feed and permeate
at the membrane surface, respectively.
The membrane distillation coefficient, C, is a function of
both the operating conditions and the membrane structure, and
depends on the MD configuration employed as well as on the
Knudsen number (Kn), which is the ratio of the mean free path of
the transported gas molecules () through the membrane pores
to the mean pore diameter of the membrane (d). In fact, Kn number determines the physical nature of flow through membrane
pores. However, since the membranes used in MD exhibit pore
size distribution, more than one mechanism may occur through
the membrane. More details about this issue are reviewed in Ref.
[1] and later well explained in various studies [8083].
8.3. Rate controlling step in MD process
Knowledge of the rate-controlling steps in each MD configuration is important in order to improve the MD systems and
to overcome the MD barriers. Either the resistance exerted by
the boundary layers or the resistance of the membrane itself
controls the MD process steps. Generally, it is admitted that
the rate limiting steps is heat transfer through boundary layers at low flow rates, while at higher flow rates the membrane
resistance becomes predominant [1,5,23,72,79,135]. Systematic
studies have been conducted in determining the process rate controlling mass transport in VMD [6,42]. It was admitted that the
heat transfer through the feed liquid phase is the controlling step
for membranes of higher permeability when only water was used
as feed. On the contrary, for membranes with small mean pore
size and low porosity, the mass transfer mechanism through the
membrane is the controlling MD step. When aqueous solution
containing volatile components such as alcohols was used, it
was found that mass transfer within the liquid phase boundary
layer dominates the mass transport process due to development
of concentration polarization [6]. It is worthy to point out that
in addition to the mass transfer resistance offered by the liquid phase, the heat transfer resistance also plays a role in the
transport process, the contribution of which may be different for
different mixtures and it is more important for highly volatile
components [42].
In AGMD and SGMD configurations, the heat transfer coefficient in the permeate side is much smaller than that of the
feed side and therefore will dominate the overall heat transfer
coefficient [4,5,139]. In other words, in the case of SGMD, the
temperature polarization is found to be localized at the permeate
side [21]. Hence small changes in the permeate flow rate (sweeping gas) will have much effects on the SGMD flux. Thus gas
flow rate is more likely to control the SGMD process especially
when non-volatile components are present in the feed solution
[4,21,22,28,69,115,164]. Otherwise, depending on the hydrodynamic conditions and the membrane module design, the rate
controlling step may switch from being mass transfer controlled

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

in the gas phase to mass transfer controlled in the liquid phase


and vice versa [132,164]. It is noteworthy to mention that despite
of the enhanced fluxes achieved while working at high air/gas
sweeping flow rates, an optimum sweep gas flow velocity is
more likely to exist [4].
The rate controlling step in MD may not only differ from one
MD configuration to another but further depend on the module
type and geometry, flow mode and operating condition, as well
as membrane morphology and process application.
8.4. Membrane fouling
In membrane separation processes, the deposition and build
up of undesirable materials on membrane surface and/or membrane pores may reduce the process permeate flux and process
efficiency. This phenomenon, known as fouling, is less studied in
MD systems. The deposited undesirable layer on the feed membrane surface may be formed by suspended particles, corrosion
products, biological growth and variety of crystalline deposits
formed as a result of scale formation fouling and thus increase
the risk of membrane pore wetting. In spite of the information
available in literature, the role of fouling in MD is not yet clearly
understood. One may expect that the role of fouling in MD may
some how differ from that exhibited in the pressure-driven membrane processes (i.e. RO, NF, UF, etc.) and completely different
from that encountered in heat exchangers. In the later systems,
fouling once happened would cause both hydraulic resistance
and heat transfer resistances. However, the penalty of increasing pressure drop is more of concern in heat exchangers than the
increase of heat transfer resistance, especially in compact heat
exchangers [146].
In MD the situation might be quit different, as foulant particles/molecules once attached to the membrane surface would
cause plugging of the membrane pore entrances causing first
some flux decay and may lead to membrane pore wetting. Moreover, due to the fact that the flow channels in MD systems are
very small, especially in hollow fiber membrane modules (i.e.
micron range), the increased deposition of the foulant species at
the membrane surface would eventually lead to an increase in the
pressure drop to levels that the hydrostatic pressure may exceed
the LEP of the feed or permeate solution into the membrane
pores.
8.4.1. Crystallization fouling
Scale formation fouling results from deposition or usually
growth of crystals on membrane surfaces during the treatment
of salt concentrated feed solutions [110]. To our knowledge,
crystal growth of salt on MD membrane surface has been first
reported in early 1980s [141]. Scale formation and deposition
at membrane surfaces may diminish the membrane hydrophobicity and cause water logging of some membrane pores. In
fact, membrane scaling is one of the major problems in seawater
desalination applications [110]. A rapid decrease of the permeate
flux associated with the precipitation of CaCO3 onto the membrane surface has been detected during water demineralization
from tap water [98]. A significant flux decline associated with
sever fouling deposits on the feed side of the membrane was also

23

reported in [142] during the concentration of NaCl solution containing dissolved organic matter by MD process. Examination
of both the membrane morphology and composition of fouling
layer showed that the membrane at the feed side was completely
covered with fouling layer containing a significant amount of
NaCl and proteins. Crystal formations on membrane surface
have been detected in DCMD configuration [64,142,143]. On
the other hand, almost no fouling was observed in an AGMD
module during a long-term run that lasted for about 2 months
[5]. It is worthy to mention here that scale formation or crystallization fouling is usually associated with aqueous solutions in
which salts that have inverse solubility characteristics (that is to
say, the solubility of which decreases with increase the temperature of solution) cause crystalline problems or scale formation.
A mathematical model was developed to predict the supersaturation profiles in the feed channel of a flat-sheet AGMD module
[110]. The model is claimed to calculate velocity and temperature profiles, as well as concentration distribution of sparingly
soluble salts such as BaSO4 , in the feed side. It was found that
for salts with inverse solubility characteristics, maximum supersaturation can occur in the bulk of the flow rather than at the
membrane surface. As a result most of membrane scaling in MD
systems can be expected with salts having a lower solubility at
lower temperatures.
8.4.2. Biological fouling
This refers to the attachment to the membrane surface of biological microorganisms such as bacteria, algae and fungi and/or
macro-organisms including mussels, barnacles, hydroids and
sapodilla worms vegetation such as seaweed. Because biological fouling is caused by one or more of the previous mentioned
living matter it is usually associated with aqueous systems where
the temperature is not too different from that in the natural
environment. A rapid flux decline due to scale formation and
biological fouling has been observed during the concentration
by DCMD of saline wastewater from the production of heparin [144]. Boiling the wastewater for about 30 min followed by
suitable filtration found to prevent their occurrence. It was also
observed a limited growth of microorganisms on the membrane
surfaces of modules previously used in different MD applications [145]. It was stated that microbial growth was significantly
affected by the MD operation conditions.
8.4.3. Particulate and corrosion fouling
Particulate fouling may generally be defined as the deposition
of solid particles in suspension onto the membrane surface. In a
fluid flowing under streamline conditions, with a concentration
gradient of a component at right angles to the direction of flow,
there is always a tendency for particulate deposition as a result
of the random motion of the molecules within the fluid system
[146]. The size of the suspended particles will have very high
influence on the deposition rate and mechanism responsible for
its transport to the membrane surface. For instance very small
particles would be expected to be subjected to Brownian diffusion and eddy diffusion, whereas larger particles because of
their mass would move under momentum forces [147]. One of
the main features of MD is that it is relayed on separation sur-

24

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

faces that are not corrodible even when they are subjected to a
highly corrosive environment, due to the fact that these surfaces
are made from polymer materials. However, when fouling is concerned, and with regards to microporous membranes involved,
one may not ignore the risk of foulant particles deposition at the
membrane surface plugging some of its pores due to corrosion
in other parts of the plant (pumps, heat exchangers, tubes/pipes,
etc.). In fact corrosion fouling may not only result in pore plugging but may even lead to damage the membrane by scratching
its surface during the movements of the corroded flakes and
chunks within the narrow flow passages.
It is worth quoting that a group of MD authors accepts the
presence of fouling in MD systems [64,142,144,145,148], while
others do not admit talking about MD fouling [51,61,62]. Most
of the performed fouling studies so far suggests that the effect
of membrane fouling if there is any, is very little compared to
other effects such as the concentration polarization, which is
even smaller than the temperature polarization effect. However
it should be emphasized here that fouling is a time dependent
process and its effect cannot be predicted.
Care must be taken when studying MD fouling effect as different membrane types may exhibit different degree of fouling,
which may depend upon the membrane hydrophobicity, membrane surface structure, initial permeability, feed solution, etc.
[61,62]. It was reported that more hydrophobic membranes may
exhibit higher degree of fouling [61,62]. Although the importance of understanding the fouling phenomena in MD has been
pointed out [1], lack of data and knowledge about fouling in MD
process still exists [148].
8.5. Long-term MD performance
Process stability during long-term exposure to different operational conditions is one of the important aspects that can convince the industrial implemtation of any new technology. MD
literature reveals that only very few studies have paid attention
to this point. It has been reported that the hydrophobicity of
the PP membrane, used during the hydrochloric acid recovery
by DCMD, was maintained and the membrane properties were
unchanged during the experiments that lasted above 1 year [133].
DCMD experiments have also been performed during more
than 4 months in Ref. [135]. As the aim of their long run was
to investigate fouling and scale formation, a normal tap water
with a preliminary conductivity of 350 S/cm was used as feed
for their DCMD module, the concentration of which was raised
after few hours to a maximum of 3500 S/cm. During the first
few weeks of their run, a flux decline was observed. Treatment
with hydrochloric acid had resulted in a complete restoration
of the original flux levels. It is worthy to mention that although
the original flux levels were restored after each acid treatment
process, flux decline seemed to be dominant on approaching the
11th week and thereafter. However, flux leveled during the last
4 weeks of their run at a value of approximately 20% lower
than the original one. Utilizing an AGMD module, two long run
experiments were performed in Ref. [5], the first with PVDF
membrane and lasted for a continuous 2 months, while the second in which PTFE membrane was employed lasted for 6 weeks.

The first 2-month duration experiment was conducted in three


stages, a month duration stage in which tap water with conductivity of 297 S/cm was used, followed by a second stage of 2
weeks period, in which sea water with an average conductivity of
about 37.6 103 S/cm was the feeding stream before switching back the to tap water in the third stage that lasted for another 2
weeks. During this 2-month run experiment, it was claimed that
neither fouling nor scaling problems were encountered, despite
the membrane wetting observed as result of electrical shutdown
occurred at the 44 days of the run. It was suspected that some
crystals formed on the membrane surface. In order to verify
their thoughts, the second run was conducted but with only tap
water as the feeding stream [5]. A steady state flux level was
approached at the seventh day of their run. In both experiments,
an initiation period of about 23 days in which an almost rapid
increase in flux was observed, followed by a slow flux decline
during the following 58 days before reaching a steady state
level. However, no explanation was given for the observed flux
variation with time. Artificial seawater was used as feed into an
AGMD module during a 10 days experiment in Ref. [24]. It was
observed that there was an increase in the flux during the first
25 h before reaching a steady state value that lasted until the end
of the run. The flux trend in this run coincided with what was
observed by the same authors in previous experiments [5] with
the only exception of time duration, which was short otherwise
the general trend of flux decline might be observed as well. One
may also expect that a similar trend for flux variation with time
might be obtained if no acid was added [135]. A flux decline by
more than 20% during 5-day continuous DCMD run carried out
with brine as feed solution have been observed [149]. Although
the authors did not investigate the reason for the observed flux
decline, they did not drop the role of dirt and fouling that might
have caused this flux decline.
The point to make here is that in all the long-term runs performed by several MD investigators, the trend of flux with time
may be, somehow, matches the ones generally observed in heat
exchanger fouling. The overall heat transfer coefficient shows
initially some enhancements as a result of the increased surface
roughness imposed by some fouling depositions. However, such
enhancements usually vanish as the deposited foulant builds up,
increasing the thermal resistance and pressure drop especially
in compact heat exchangers.
A new research paper appeared recently [150], in which the
results of what seems the longest run research (i.e. over 3 years)
on a DCMD module used for the production of pure water have
been presented. Interestingly the used hydrophobic capillary PP
membranes were found to be thermally stable and good separation characteristics were maintained through out the whole
period of their investigation.
As can be concluded from the exposed results, more studies
must be conducted on this fact to help understanding why in
some cases MD flux decline is detected while in others not.
8.6. Energy and maintenance cost evaluation in MD
The commercial technology of MD is in an early stage,
and the logic first step is to produce working, reliable mem-

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

brane modules [107]. This may explain the reason of scattered


available sources on cost estimation within the MD literature,
especially in view of the wide variety of areas in which MD is
applicable. Detailed approximation cost analysis of a scaled up
plant (of 5000 kg/h of potable water production) made of DCMD
hollow fibers module that providing total surface area of 800 m2
and utilizing a heat recovery system was reported in Ref. [151].
It was stated that the production costs are very sensitive to the
feed temperature, and at a temperature of 80 C the plant production costs could be competitive to that of RO. Moreover, the
costs for a small pilot plant of production capacity of 50 kg/h
(500 kg/day), utilizing solar energy suitable for installation in
arid/rural area was estimated. The cost estimation for this pilot
plant was somewhat more costly than a RO desalination plant
of the same capacity [151]. The feasibility of solar powered
MD (SPMD) pilot plant for the supply of domestic drinking
water in the arid/rural regions of Australia was examined [106].
With the aid of a computer simulation programme, the optimum
heat recovery system was selected. It was found that for a pilot
plant of 50 kg/h capacity, the heat exchangers are the most costly
items. In order to reduce the capital cost of heat exchangers, their
optimized design shifted towards using more heat recovery (in
the range 6080%) and less solar collector area. However, no
production costs were provided in their study, but only capital
itemized costs.
Realizing the fact that operating costs represents a large fraction of the production cost and the thermal energy consumption
is very much sensitive to the feed temperature [22,151], utilizing a heat recovery system should be an essential part of any
efficient MD plant. The heat supplied for vaporization to the
feed streams, to a much extent, should be recovered from the
permeate streams to minimize the energy requirements. This
would not be achieved without the utilization of a heat recovery
function (i.e. heat exchanger), which means increasing capital investment costs but benefiting operating costs. Among the
MD configurations, VMD is more likely to be the most energy
efficient configuration, due to its low heat loss by conduction
through membrane and as result of its high flux rates. On the
other hand, loss of driving force as result of imposed temperature polarization, heat losses through the membrane and
to the environment should be minimized. While the temperature polarization effect can be diminished by working under
turbulent flow regime and with the aid of turbulence promoters (i.e. good module design), the heat loss by conduction
through the membrane can be minimized by increasing pores
size and membrane porosity as well as by employing thicker
membranes. However, the membrane thickness increases the
resistance to mass transport due to the increased path length.
Proper isolation of membrane module, its piping and other
accessories can reduce or even limit heat loss to environment.
It was reported that the amount of heat loss by conduction
in the DCMD configuration, which was the less heat efficient
among the MD configurations, can be negligible at high feed
temperatures in comparison to the effective heat in evaporation
[22]. Therefore, the energy consumption per unit of distillate
may be reduced appreciably at high operating feed temperatures.

25

At present, one of the barriers for MD implementation is


the high cost of commercial modules. In fact, there is a lack
of commercially available MD units because most researchers
use modules designed for other membrane operations. However, prior to any cost estimation, one may judge that in view
of their very low operating pressures, allowing for thinner
piping and fewer working problems, the capital and maintenance expenses for MD should be lower in comparison to
pressure-driven membrane separation processes like UF and RO.
Nonetheless, for potential commercial application of MD technology, much larger membrane modules having the required
effective surface area should be investigated [107]. Assuming
fully developed MD technology, the total production cost for
purified water production rate at 3800 m3 /day has been estimated
[149]. The assumed DCMD desalination plant with a heat recovery of 30% was found to be feasible, with total production costs
of $2.97/1000 gal (=$0.782 m3 ), which was quite feasible in
comparison to that of RO of $4.77/1000 gal (=$1.26 m3 ) of the
same capacity. The energy requirements related to integrated
systems of RO/MD, in which MD operates on the RO brine,
and NF/RO/MD, in which NF is introduced for water pretreatment were determined [97]. Their analysis showed that despite
of the higher energy consumption associated with the coupling
systems with respect to operating RO alone, the overall performance was appreciable. The integrated MD systems can be very
attractive alternative to RO if the thermal energy is available
[97].

9. Conclusions
For any new technology to be industrially accepted and implemented there is a need to verify all of its claimed benefits as well
as to show its superiority over other already well established
technologies. This has first to be achieved on the laboratory
scale prior to pilot plant scale to finally reach industry. Surprisingly, for a process such as MD that for long has been attractive
and claimed to be a promising separation technique, the first step
(laboratory scale) has not yet been fulfilled. Study and evaluation
of the majority of the MD literature reveal that the process still
suffers from lack and inconsistency of information about some
of its major aspects. On the ground of acquired knowledge and
gathered information on MD process, a framework for better
understanding of the MD process was presented in this study.
This framework is an attempt to highlight those aspects of the
MD process that are far from being understood and also serves
as a preliminary guide showing the sequences of the different
steps that should be followed to fulfill the requirements for MD
industrial application.

Acknowledgement
This work is financially supported by National Basic
Research Program of China (973 Program) No. 2003CB615700
and State Key Project for Fundamental Research Development
(Project973 of China, No. 2004CB719700).

26

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

References
[1] K.W. Lawson, D.R. Lloyd, Membrane distillation, J. Membr. Sci. 124
(1997) 125.
[2] J.I. Mengual, L. Pena, Membrane distillation, Colloid Interf. Sci. 1 (1997)
1729.
[3] R.W. Schofield, A.G. Fane, C.J.D. Fell, Heat and mass transfer in membrane distillation, J. Membr. Sci. 33 (1987) 299313.
[4] M. Khayet, M.P. Godino, J.I. Mengual, Theory and experiments on sweeping gas membrane distillation, J. Membr. Sci. 165 (2000) 261272.
[5] F.A. Banat, J. Simandl, Theoretical and experimental study in membrane
distillation, Desalination 95 (1994) 3952.
[6] S. Bandini, C. Gostoli, G.C. Sarti, Separation efficiency in vacuum membrane distillation, J. Membr. Sci. 73 (1992) 217229.
[7] S. Kimura, S.I. Nakao, S.I. Shimatani, Transport phenomena in membrane
distillation, J. Membr. Sci. 33 (3) (1987) 285298.
[8] B.R. Bodell, Silicone rubber vapor diffusion in saline water distillation,
United States Patent no. 285,032 (1963).
[9] M.E. Findley, Vaporization through porous membranes, Ind. Eng. Chem.
Process Des. Dev. 6 (1967) 226230.
[10] D.W. Gore, Gore-Tex membrane distillation, in: Proceedings of the 10th
Annual Convention of the Water Supply Improvement Assoc, Hononulu,
USA, 1982.
[11] S.I. Anderson, N. Kjellander, B. Rodesjo, Design and field tests of a
new membrane distillation desalination process, Desalination 56 (1985)
345354.
[12] K. Schnieder, T.J. van Gassel, Membrane distillation, Chem. Eng. Technol. 56 (1984) 514521.
[13] C. Bier, U. Plantikow, Solar powered desalination by membrane distillation, in: IDA World Congress on Desalination and Water Science, Abu
Dhabi, 1995, pp. 397410.
[14] J. Koschikowski, M. Wieghaus, M. Rommel, Solar thermal-driven desalination plants based on membrane distillation, Desalination 156 (2003)
295304.
[15] J.I. Mengual, K. Khayet, M.P. Godino, Heat and mass transfer in vacuum
membrane distillation, Int. J. Heat Mass Trans. 47 (2004) 865875.
[16] J. Phattaranawik, R. Jiraratananon, A.G. Fane, Heat transport and membrane distillation coefficients in direct contact membrane distillation, J.
Membr. Sci. 212 (2003) 177193.
[17] M. Khayet, T. Matsuura, Preparation and characterization of polyvinylidene fluoride membranes for membrane distillation, Ind. Eng. Chem. Res.
40 (2001) 57105718.
[18] M. Khayet, T. Matsuura, Application of surface modifying macromolecules for the preparation of membranes for membrane distillation,
Desalination 158 (2003) 5156.
[19] C. Feng, B. Shi, G. Li, Y. Wu, Preparation and properties of microporous
membrane from poly (vinylidene fluoride-co-tetrafluoro ethylene) (F2.4)
for membrane distillation, J. Membr. Sci. 237 (2004) 1524.
[20] J. Phattaranawik, R. Jiraratananon, Direct contact membrane distillation: effect of mass transfer on heat transfer, J. Membr. Sci. 188 (2001)
137143.
[21] M. Khayet, M.P. Godino, J.I. Mengual, Thermal boundary layers in
sweeping gas membrane distillation processes, Am. Inst. Chem. Eng.
J. (AIChE J.) 48 (7) (2002) 14881497.
[22] M. Khayet, M.P. Godino, J.I. Mengual, Possibility of nuclear desalination through various membrane distillation configurations: a comparative
study, Int. J. Nucl. Desalinat. 1 (1) (2003) 3046.
[23] K.W. Lawson, D.R. Lloyd, Membrane distillation. I. Module design and
performance evaluation using vacuum membrane distillation, J. Membr.
Sci. 120 (1996) 111121.
[24] F.A. Banat, J. Simandl, Desalination by membrane distillation: a parametric study, Sep. Sci. Technol. 33 (2) (1998) 201226.
[25] M.A. Izquierdo-Gil, M.C. Garcia-Payo, C. Fernandez-Pineda, Air gap
membrane distillation of sucrose aqueous solutions, J. Membr. Sci. 155
(2) (1999) 291307.
[26] C.A. Rivier, M.C. Garcia-Payo, I.W. Marison, U.V. Stockar, Separation
of binary mixtures by thermostatic sweeping gas membrane distillation.
I. Theory and simulations, J. Membr. Sci. 201 (2002) 116.

[27] M.C. Garcia-Payo, C.A. Rivier, I.W. Marison, U.V. Stockar, Separation
of binary mixtures by thermostatic sweeping gas membrane distillation.
II. Experimental results with aqueous formic acid solutions, J. Membr.
Sci. 198 (2002) 197210.
[28] M. Khayet, M.P. Godino, J.I. Mengual, Theoretical and experimentalstudies on desalination using the sweeping gas membrane distillation,
Desalination 157 (2003) 297305.
[29] M. Khayet, J.I. Mengual, G. Zakrzewska-Trznadel, Direct contact
membrane distillation for nuclear desalination. Part II. Experiments
with radioactive solutions, Int. J. Nucl. Desalinat. (IJND) 56 (2006)
5673.
[30] M. Sudoh, K. Takuwa, H. Iizuka, K. Nagamatsuya, Effects of thermal and
concentration boundary layers on vapor permeation in membrane distillation of aqueous lithium bromide solution, J. Membr. Sci. 131 (1997)
17.
[31] E. Drioli, V. Calabr`o, Y. Wu, Microporous membranes in membrane distillation, Pure Appl. Chem. 58 (12) (1986) 16571662.
[32] P.P. Zolotarev, V.V. Ugrosov, I.B. Volkina, V.N. Nikulin, Treatment of
waste-water for removing heavy-metals by membrane distillation, J. Hazard. Mater. 37 (1) (1994) 7782.
[33] S.H. Duan, A. Ito, A. Ohkawa, Removal of trichloroethylene from water
by aeration, pervaporation and membrane distillation, J. Chem. Eng. Jpn.
34 (8) (2001) 10691073.
[34] F.A. Banat, J. Simandl, Removal of benzene traces from contaminated
water by vacuum membrane distillation, Chem. Eng. Sci. 51 (8) (1996)
12571265.
[35] G.C. Sarti, C. Gostoli, S. Bandini, Extraction of organic-compounds from
aqueous streams by vacuum membrane distillation, J. Membr. Sci. 80
(1993) 2133.
[36] N. Qureshi, M.M. Meagher, R.W. Hutkins, Recovery of 2,3-butanediol
by vacuum membrane distillation, Sep. Sci. Technol. 29 (13) (1994)
17331748.
[37] F.A. Banat, M. Al-Shannag, Recovery of dilute acetonebutanolethanol
(ABE) solvents from aqueous solutions via membrane distillation, Bioprocess Eng. 23 (6) (2000) 643649.
[38] F.A. Banat, J. Simandl, Membrane distillation for propane removal from
aqueous streams, J. Chem. Technol. Biotechnol. 75 (2) (2000) 168178,
AGMD.
[39] M.C. Garcia-Payo, M.A. Izquierdo-Gil, C. Fernandez-Pineda, Air gap
membrane distillation of aqueous alcohol solutions, J. Membr. Sci. 169
(2000) 6180.
[40] F.A. Banat, J. Simandl, Membrane distillation for dilute ethanol separation from aqueous streams, J. Membr. Sci. 163 (1999) 333348.
[41] S. Bandini, A. Saavedra, G.C. Sarti, Vacuum membrane distillation:
experiments and modeling, AIChE J. 43 (2) (1997) 398408.
[42] S. Bandini, G.C. Sarti, Heat and mass transfer resistances in vacuum membrane distillation per drop, AIChE J. 45 (7) (1999) 1422
1433.
[43] S. Nene, S. Kaur, K. Sumod, B. Joshi, K.S.M.S. Raghavarao, Membrane
distillation for the concentration of raw cane-sugar syrup and membrane
clarified sugarcane juice, Desalination 147 (2002) 157160.
[44] C. Varming, M.L. Andersen, L. Poll, Influence of thermal treatment on
black currant (Ribes nigrum L.) juice aroma, J. Agric. Food Chem. 52
(25) (2004) 76287636.
[45] M. Gryta, K. Karakulski, The application of membrane distillation for the
concentration of oilwater emulsions, Desalination 121 (1999) 2329.
[46] V. Calabr`o, B.L. Jiao, E. Drioli, Theoretical and experimental study on
membrane distillation in the concentration of orange juice, Ind. Eng.
Chem. Res. 33 (71) (1994) 18031808.
[47] F. Lagan`a, G. Barbieri, E. Drioli, Direct contact membrane distillation:
modelling and concentration experiments, J. Membr. Sci. 166 (2000)
111.
[48] K. Sakai, T. Muroi, K. Ozawa, S. Takesawa, M. Tamura, T. Nakaue,
Extraction of solute-free water from blood by membrane distillation,
Trans. Am. Soc. Artif. Intern. Organs 32 (1986) 397400.
[49] K. Sakai, T. Koyano, T. Muroi, M. Tamura, Effects of temperature and
concentration polarization on water vapor permeability for blood in membrane distillation, Chem. Eng. J. 38 (1988) B33B39.

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429


[50] K. Sakai, T.K. Ano, T. Murai, et al., Effect of temperature polarization
on water vapor permeability for blood in membrane distillation, Chem.
Eng. Jpn. 38 (1988) 833.
[51] J.M. Ortiz de Zarate, C. Rincon, J.I. Mengual, Concentration of bovine
serum albumin aqueous solutions by membrane distillation, Sep. Sci.
Technol. 33 (3) (1998) 283296.
[52] A. Capuano, B. Menoli, V.E. Andreucci, A. Criscuoli, E. Drioli, Membrane distillation of human plasma ultrafiltrate and its theoretical applications to haemodialysis techniques, Int. J. Artif. Organs 23 (7) (2000)
415422.
[53] R.R. Nigmatullui, A.F. Burban, A.F. Melnik, M.T. Bryk, V.V. Kondratyuk,
Concentration of solutions of thermolabile vitamins by membrane distillation, Russian J. Appl. Chem. 66 (6) (1993) 10701073.
[54] Y. Wu, Y. Kong, J. Liu, J. Zhang, J. Xu, An experimental study on
membrane distillation-crystallization for treating waste water in taurine
production, Desalination 80 (1991) 235242.
[55] V. Calabr`o, E. Drioli, F. Matera, Membrane distillation in the textile
wastewater treatment, Desalination 83 (1991) 209224.
[56] F. Banat, S. Al-Asheh, M. Qtaishat, Treatment of waters colored with
methylene blue dye by vacuum membrane distillation, Desalination 174
(2005) 8796.
[57] A.G. Chmielewski, G. Zakrzewska-Trznadel, N.R. Miljevic, W.A. Van
Hook, Membrane distillation employed from separation of water isotopic
compounds, Sep. Sci. Technol. 30 (7/9) (1995) 16531667.
[58] M. Tomaszewska, Concentration of the extraction of fluid from sulfuric
acid treatment of phosphogypsum by membrane distillation, J. Membr.
Sci. 78 (1993) 277282.
[59] M. Tomaszewska, M. Gryta, A.W. Morawski, Study on the concentration
of acids by membrane distillation, J. Membr. Sci. 102 (1995) 113122.
[60] J.J. Tang, K.G. Zhou, F.G. Zhao, R.X. Li, Q.X. Zhang, Hydrochloric
acid recovery from rare earth chloride solutions by vacuum membrane
distillation (1) Study on the possibility, J. Rare Earths 21 (2003) 78
82.
[61] M. Khayet, A. Velazquez, J.I. Mengual, Direct contact membrane distillation of humic acid solutions, J. Membr. Sci. 240 (2004) 123128.
[62] M. Khayet, J.I. Mengual, Effect of salt concentration during the treatment
of humic acid solutions by membrane distillation, Desalination 168 (2004)
373381.
[63] E. Curcio, A. Criscuoli, E. Drioli, Membrane crystallizers, Ind. Eng.
Chem. Res. 40 (2001) 26792684.
[64] C.M. Tun, A.G. Fane, J.T. Matheickal, R. Sheikholeslami, Membrane
distillation crystallization of concentrated salts-flux and crystal formation,
J. Membr. Sci. 257 (2005) 144155.
[65] E. Drioli, A. Criscuoli, E. Curcio, Integrated membrane operation for
seawater desalination, Desalination 147 (2002) 7781.
[66] S. Bouguecha, M. Dhahbi, Fluidized bed crystallizer and air gap membrane distillation as a solution to geothermal water desalination, Desalination 152 (2002) 237244.
[67] M.C. Garca-Payo, M.A. Izquierdo-Gil, C. Fernandez-Pineda, Wetting
study of hydrophobic membranes via liquid entry pressure measurements with aqueous alcohol solutions, J. Colloids Interf. Sci. 230 (2000)
420431.
[68] P.K. Weyl, Recovery of demineralized water from saline waters, United
States Patent no. 3,340,186 (1967).
[69] M. Khayet, M.P. Godino, J.I. Mengual, Study of asymmetric polarization
in direct contact membrane distillation, Sep. Sci. Technol. 39 (1) (2004)
125147.
[70] L. Martinez-Diez, M.I. Vazquez-Gonzalez, Effect of polarization on mass
transport through hydrophobic porous membranes, J. Ind. Eng. Chem.
Res. 37 (1998) 41284135.
[71] L. Martinez-Diez, M.I. Vazquez-Gonzalez, Temperature and concentration polarization in membrane distillation of aqueous salt solutions, J.
Membr. Sci. 156 (1999) 265273.
[72] R.W. Schofield, A.G. Fane, C.J.D. Fell, R. Macoun, Factors affecting flux
in membrane distillation, Desalination 77 (1/3) (1990) 279294.
[73] C. Zhu, G.L. Liu, C.S. Cheung, C.W. Leung, Z.C. Zhu, Ultrasonic stimulation on enhancement of air gap membrane distillation, J. Membr. Sci.
161 (1999) 8593.

27

[74] C. Zhu, G. Liu, Modeling of ultrasonic enhancement on membrane distillation, J. Membr. Sci. 176 (2000) 3141.
[75] T.Y. Cath, V.D. Adama, A.E. Childress, Experimental study of distillation using direct contact membrane distillation: a new approach to flux
enhancement, J. Membr. Sci. 228 (2004) 516.
[76] L. Martinez-Diez, M.I. Vazquez-Gonzalez, F.J. Florido-Diaz, Study of
membrane distillation using channel spacers, J. Membr. Sci. 144 (1/2)
(1998) 4556.
[77] J. Phattaranawik, R. Jiraratananon, A.G. Fane, C. Halim, Mass flux
enhancement using spacer filled channels in direct contact membrane
distillation, J. Membr. Sci. 187 (2001) 193201.
[78] R.W. Schofield, A.G. Fane, C.J.D. Fell, Gas and vapor transport through
microporous membrane. I. KnudsenPoiseuille transition, J. Membr. Sci.
53 (1990) 159171.
[79] R.W. Schofield, A.G. Fane, C.J.D. Fell, Gas and vapor transport through
microporous membrane. II, J. Membr. Sci. 53 (1990) 173185.
[80] J. Phattaranawik, R. Jiraratananon, A.G. Fane, Effect of pore size distribution and air flux on mass transport in direct membrane distillation, J.
Membr. Sci. 215 (2003) 7585.
[81] M. Khayet, K.C. Khulbe, T. Matsuura, Characterization of membrane
distillation by atomic force microscopy and estimation of their water
vapor transfer coefficients in vacuum membrane distillation process, J.
Membr. Sci. 238 (2004) 199211.
[82] V.V. Ugrozov, I.B. Elkina, Mathematical modeling of influence of porous
structure of membrane on its vapor-conductivity in the process of membrane distillation, Desalination 147 (2002) 167171.
[83] L. Martinez-Diez, F.J. Florido-Diaz, A. Hernandez, P. Pradanos, Characterization of three hydrophobic porous membranes used in membrane
distillation, J. Membr. Sci. 203 (2002) 1527.
[84] M. Khayet, T. Matsuura, Pervaporation and vacuum membrane distillation processes: modeling and experiments, AIChE J. 50 (8) (2004)
16971712.
[85] Y. Fujii, S. Kigoshi, H. Iwatani, M. Aoyama, Selectivity and characteristics of direct contact membrane distillation type experiments. I. Permeability and selectivity through dried hydrophobic fine porous membranes,
J. Membr. Sci. 72 (1992) 5372.
[86] J. Li, Z. Xu, Z. Liu, W. Yuan, H. Xiang, S. Wang, Y. Xu, Micropous
polypropylene and polyethylene hollow fiber membranes. Part 3. Experimental studies on membrane distillation for desalination, Desalination
155 (2003) 153156.
[87] J.M. Ortiz de Zarate, L. Pena, J.I. Mengual, Characterization of membrane
distillation membranes prepared by phase inversion, Desalination 100
(1996) 139148.
[88] M. Tomaszewska, Preparation and properties of flat-sheet membranes
from poly(vinylidene) for membrane distillation, Desalination 104 (1996)
111.
[89] M. Khayet, J.I. Mengual, T. Matsuura, Porous hydrophobic/hydrophilic
composite membranes: application in desalination using direct contact
membrane distillation, J. Membr. Sci. 252 (2005) 101113.
[90] Y. Wu, Y. Kong, X. Lin, W. Liu, J. Xu, Surface-modified hydrophilic
membranes in membrane distillation, J. Membr. Sci. 72 (1992) 189196.
[91] K.W. Lawson, D.R. Lloyd, Membrane distillation. II. Direct contact membrane distillation, J. Membr. Sci. 120 (1996) 123133.
[92] P.J. Foster, A. Burgoyne, M.M. Vahdati, Improved process topology for
membrane distillation, Sep. Purif. Technol. 21 (2001) 205217.
[93] M. Khayet, T. Matsuura, J.I. Mengual, M. Qtaishat, Design of novel
direct contact membrane distillation membranes, Desalination 192 (2006)
105111.
[94] B. Li, K.K. Sirkar, Novel membrane and device for direct contact membrane distillation-based desalination process, Ind. Eng. Chem. Res. 43
(2004) 53005309.
[95] B. Li, K.K. Sirkar, Novel membrane and device for vacuum membrane
distillation-based desalination process, J. Membr. Sci. 257 (2005) 60
75.
[96] C. Cabassud, D. Wirth, Membrane distillation for water desalination: how
to chose an appropriate membrane, Desalination 157 (2003) 307314.
[97] A. Criscuoli, E. Drioli, Energetic and exergetic analisis of an integrated
membrane desalination, Desalination 124 (1999) 243249.

28

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429

[98] K. Karakulski, M. Gryta, A. Morawski, Membrane processes for portable


water quality improvement, Desalination 145 (2002) 315319.
[99] Z. Wang, F. Zheng, S. Wang, Experimental study of membrane distillation
with brine circulated in the cold side, J. Membr. Sci. 183 (2001) 171179.
[100] T.Y. Cath, D. Adams, A.E. Childress, Membrane contactor processes for
wastewater reclamation in space. II. Combined direct osmosis, osmotic
distillation, and membrane distillation for treatment of metabolic wastewater, J. Membr. Sci. 257 (2005) 111119.
[101] V. Calabro, G. Pantano, M. Kang, R. Molinari, E. Drioli, Experimental study on integrated membrane proceses in the treatment of solutions
simulating textile effluents: energy and exergy analisis, Desaliantion 78
(1990) 257277.
[102] M.C. Andres, J. Doria, M. Khayet, L. Pena, J.I. Mengual, Coupling of
a membrane distillation module to a multieffect distiller for pure water
production, Desalination 115 (1998) 7181.
[103] G. Zakrzewska-Trznadel, M. Harasimowicz, A.G. Chmielewski, Membrane processes in nuclear technology-application for liquid radioactive
waste treatment, Sep. Purif. Technol. 22/23 (2001) 617625.
[104] M. Khayet, J.I. Mengual, G. Zakrzewska-Trznadel, Direct contact membrane distillation for nuclear desalination. Part I. Review of membranes
used in membrane distillation and methods for their characterization, Int.
J. Nucl. Desalinat. (IJND) 1 (4) (2005) 435449.
[105] G. Zakrewska-Trznadel, M. Harasimowicz, A.G. Chmielewski, Concentration of radioactive components in liquid low-level radioactive waste
by membrane distillation, J. Membr. Sci. 163 (1999) 257264.
[106] P.A. Hogan, Sudjito, A.G. Fane, G.L. Morrison, Desalination by solar
heated membrane distillation, Desalination 81 (1991) 8190.
[107] J. Walton, H. Lu, C. Tumer, S. Solis, H. Hein, Solar and waste heat
desalination by membrane distillation, Desalination and Water Purification Research and Development Program Report No. 81, 2004.
[108] M. Banat, R. Jumah, M. Garaibeh, Exploitation of solar energy collected
by solar stills from desalination by membrane distillation, Renew. Energy
25 (2002) 293305.
[109] Z. Ding, L. Liu, M.S. EL-Bourawi, R. Ma, Analysis of a solar-powered
membrane distillation system, Desalination 172 (2005) 2740.
[110] M.N. Chernyshov, G.W. Meindersma, A.B. De-Haan, Modelling temperature and salt concentration distribution in membrane distillation feed
channel, Desalination 157 (2003) 315324.
[111] S.P. Agashichev, D.V. Falalejev, Modeling temperature polarization phenomena for longitudinal shell-side flow in membrane distillation process,
Desalination 108 (1996) 99103.
[112] S.P. Agashichev, A.V. Sivakov, Modeling and calculation of temperatureconcentration polarization in the membrane distillation process (MD), J.
Desalinat. 93 (1993) 245258.
[113] F. Banat, F.A. Al-Rub, K. Bani-Melhem, Desalination by vacuum membrane distillation: sensitivity analysis, Sep. Purif. Technol. 33 (2003)
7587.
[114] A.M. Alkaibi, N. Lior, Transport analysis of air-gap membrane distillation, J. Membr. Sci. 255 (2005) 239253.
[115] M. Khayet, M.P. Godino, J.I. Mengual, Nature of flow on sweeping gas
membrane distillation, J. Membr. Sci. 170 (2000) 243255.
[116] S. Bouguccha, R. Chouilkh, M. Dhahbi, Numerical study of the coupled
heat and mass transfer in membrane distillation, Desalination 152 (2002)
245252.
[117] M. Khayet, A. Velazquez, J.I. Mengual, Modelling mass transport through
a porous partition: Effect of pore size distribution, J. Non-Equil. Thermodyn. 29 (2004) 279299.
[118] M. Gryta, M. Tomaszewska, A.W. Morawski, Membrane distillation with
laminar flow, Sep. Purif. Technol. 11 (2) (1997) 93101.
[119] M. Gryta, M. Tomaszewska, Heat transport in the membrane distillation
process, J. Membr. Sci. 144 (1998) 211222.
[120] L. Pena, M.P. Godino, J.I. Mengual, A method to evaluate the net membrane distillation coefficient, J. Membr. Sci. 143 (1998) 219233.
[121] P. Godino, L. Pena, J.I. Mengual, Membrane distillation: theory and experiments, J. Membr. Sci. 121 (1996) 8393.
[122] A.O. Imdakm, T. Matsuura, A Monte Carlo simulation model for membrane distillation processes: direct contact (MD), J. Membr. Sci. 237
(2004) 5159.

[123] A.O. Imdakm, T. Matsuura, Simulation of heat and mass transfer in direct
contact membrane distillation (MD): the effect of membrane physical
properties, J. Membr. Sci. 262 (2005) 117128.
[124] C. Ferandez-Pineda, M.A. Izquierdo-Gil, M.C. Garcia-Payo, Gas permeation and direct contact membrane distillation experiments and their
analysis using different models, J. Membr. Sci. 198 (2002) 3349.
[125] L. Martinez-Diez, F.J. Florido-Diaz, A. Hernandez, P. Pradanos, Estimation of vapor transfer coefficient of hydrophobic porous membranes
for applications in membrane distillation, Sep. Purif. Technol. 33 (2003)
4555.
[126] Z. Ding, L. Liu, R. Ma, Study on the effect of flow maldistribution on
the performance of hollow fiber module used in membrane distillation, J.
Membr. Sci. 215 (1/2) (2003) 1123.
[127] X. Zhang, W. Zhang, X. Hao, H. Zhang, Z. Zhang, J. Zhang, Mathematical
model of gas permeation through PTFE porous membrane and the effect
of membrane pore structure, Chinese J. Chem. Eng. 11 (4) (2003) 383
387.
[128] Z. Ding, R. Ma, A.G. Fane, A new model for mass transfer in direct
contact membrane distillation, Desalination 151 (2002) 217227.
[129] L. Martinez-Diez, M.I. Vazquez-Gonzalez, A method to evaluate coefficients affecting flux in membrane distillation, J. Membr. Sci. 173 (2000)
225234.
[130] J.M. Rodriquez-Maroto, L. Martinez-Diez, Bulk and measured temperatures in direct contact membrane distillation, J. Membr. Sci. 250 (2005)
141149.
[131] A.M. Alkaibi, N. Lior, Membrane-distillation desalination: status and
potential, Desalination 171 (2004) 111131.
[132] C.H. Lee, W.H. Hong, Effect of operating variables on flux and selectivity
in sweep gas membrane distillation for dilute aqueous isopropanol, J.
Membr. Sci. 188 (2001) 7986.
[133] M. Tomaszewska, M. Gryta, A.W. Morawski, The influence of salt in
solution on hydrochloric acid recovery by membrane distillation, Sep.
Purif. Technol. 14 (1998) 183188.
[134] E. Drioli, Y. Wu, V. Calabro, Membrane distillation in the treatment of
aqueous solutions, J. Membr. Sci. 33 (1987) 277284.
[135] K. Schneider, W. Holz, R. Wollbeck, S. Ripperger, Membranes and modules for transmembrane distillation, J. Membr. Sci. 39 (1988) 2542.
[136] M. Khayet, T. Matsuura, J.I. Mengual, Porous hydrophobic/hydrophilic
composite membranes: estimation of the hydrophobic-layer thickness, J.
Membr. Sci. 266 (2005) 6879.
[137] K.W. Lawson, M.S. Hall, D.R. Lloyd, Compaction of microporous membranes used in membrane distillation. I. Effect on gas permeability, J.
Membr. Sci. 101 (1995) 99108.
[138] M. Khayet, M.P. Godino, J.I. Mengual, Modelling transport mechanism
through a porous partition, J. Non-Equil. Thermodyn. 26 (1/14) (2001).
[139] J. Zheng, Y. Xu, Z. Xu, Flow distribution in a randomly packed hollow
fiber membrane module, J. Membr. Sci. 211 (2003) 263269.
[140] J. Zheng, Z. Xu, J. Li, S. Wang, Y. Xu, Influence of random arrangement
of hollow fiber membranes on shell side mass transfer performance: a
novel model prediction, J. Membr. Sci. 236 (2004) 145151.
[141] D.Y. Cheng, S.J. Wiersma, Apparatus and method for thermal membrane
distillation, US Patent No. 4,419,187 (1983).
[142] M. Gryta, M. Tomaszewska, J. Grzechulska, A.W. Morawski, Membrane
distillation of NaCl solution containing natural organic matter, J. Membr.
Sci. 181 (2001) 279287.
[143] E. Drioli, Y. Wu, Membrane distillation: an experimental study, Desalination 53 (1985) 339346.
[144] M. Gryta, Concentration of saline wastewater from the production of
heparine, Desalination 129 (2000) 3544.
[145] M. Gryta, The assessment of microorganism growth in the membrane
distillation system, Desalination 142 (2002) 7988.
[146] M.S. EL-Bourawi, Performance of Polymer Film Compact Heat
Exchanger Under Fouling Conditions, MPhil. Thesis, University of Newcastle upon Tyne, 1999.
[147] T.R. Bott, R.A. Walker, Fouling in heat transfer equipment, Chem. Eng.
251 (1971) 391395.
[148] S. Srisurichan, R. Jiraratananon, A.G. Fane, Humic acid fouling in the
membrane distillation process, Desalination 174 (2005) 6372.

M.S. El-Bourawi et al. / Journal of Membrane Science 285 (2006) 429


[149] K.K. Sirkar, B. Li, Novel membrane and device for direct contact membrane distillation-based desalination process: Phase II, Desalination and
Water Purification Research and Development Program Final Report No.
96, July 2003.
[150] M. Gryta, Long-term performance of membrane distillation distillation
process, J. Membr. Sci. 265 (2005) 153159.
[151] A.G. Fane, R.W. Schofield, G.J.D. Fell, The efficient use of energy in
membrane distillation, Desalination 64 (1987) 231243.
[152] L. Martinez-Diez, F.J. Florido-Diaz, M.I. Vazquez-Gonzalez, Study of
evaporation efficiency in membrane distillation, J. Membr. Sci. 126 (1999)
193198.
[153] M.P. Godino, L. Pena, C. Rincon, J.I. Mengual, Water production from
brines by membrane distillation, Desalination 108 (1/3) (1997) 9197.
[154] C. Rincon, M.J. Ortiz de Zarate, J.I. Mengual, Separation of water and
glycols by direct contact membrane distillation, J. Membr. Sci. 158 (1999)
155165.
[155] L. Martinez, Comparison of membrane distillation performance using
different feeds, Desalination 168 (2004) 359365.
[156] Z. Ding, W. Liu, G. Zhang, R. Ma, Comparative study on direct contact and air gap membrane distillation, Mod. Chem. Ind. 22 (2002)
2629.
[157] C. Gostoli, G.C. Sarti, Law temperature distillation through hydrophobic
membranes, J. Sep. Sci. Technol. 27 (2/3) (1987) 855872.
[158] M.A. Izquierdo-Gil, J. Abildskov, G. Jonsson, The use of VMD data/
model to test different thermodynamic models for vaporliquid equilibrium, J. Membr. Sci. 239 (2004) 227241.

29

[159] L. Martinez-Diez, F.J. Florido-Diaz, Desalination of brines by membrane


distillation, Desalination 137 (2001) 267273.
[160] D. Wirth, C. Cabassud, Water desalination using membrane distillation:
comparison between inside/out and outside/in permeation, Desalination
147 (2002) 139145.
[161] J. Kim, S.E. Park, T.S. Kim, D.Y. Jeong, K.H. Ko, Isotopic water separation using AGMD and VEMD, Nukleonika 49 (4) (2004) 137142.
[162] L. Basini, G.D. Angelo, M. Gobbi, G.C. Sarti, C. Gostoli, A Desalination
process through sweeping gas membrane distillation, Desalination 64
(1987) 245.
[163] H. Mahmud, A. Kumar, R.M. Narbiatz, T. Matsuura, A study of
mass transfer in the membrane air-stripping process using microporous
polypropylene hollow fibers, J. Membr. Sci. 179 (2000) 2941.
[164] C. Boi, S. Bandini, G.C. Sarti, Pollutants removal from wastewater
through membrane distillation, Desalination 183 (2005) 383394.
[165] M.N. Chernyshov, G.W. Meindersma, A.B. de Haan, Comparison of
spacers for temperature polarization reduction in air gap membrane distillation, Desalination 183 (2005) 363374.
[166] M. Tomaszewska, M. Gryta, A.W. Morawski, Mass transfer of HCL and
H2 O across the hydrophobic membrane during membrane distillation, J.
Membr. Sci. 166 (2000) 149157.
[167] P. Peng, A.G. Fane, X. Li, Desalination by membrane distillation adopting
a hydrophilic membrane, Desalination 173 (2005) 4554.
[168] A. Bottino, G. Capannelli, A. Comite, Novel porous poly (vinylidene
fluoride) membranes for membrane distillation, Desalination 183 (2005)
375382.

Potrebbero piacerti anche