ArticlePDF Available

Enhanced Performance of Membrane Distillation Using Surface Heating Process

Authors:

Abstract and Figures

Membrane distillation (MD) is a thermally driven desalination process that has excellent application prospects in seawater desalination or hypersaline wastewater treatment, while severe temperature polarization (TP) and the resulting relatively high energy consumption have become principal challenges limiting the commercial application of MD. Therefore, the design of novel systems to overcome the shortage of conventional MD requires urgent attention. Here, we developed three surface heating vacuum membrane distillation systems, namely, SHVMD-1, SHVMD-2, and SHVMD-3, according to the different positions of the thermal conducting layer in the cell. The distillate flux, TP, and energy performance of these systems under different operating conditions were investigated. All three systems showed stable performance, with a salt rejection >99.98% for 35 g/L NaCl, and the highest flux was close to 9 L/m2·h. The temperature polarization coefficients were higher than unity in SHVMD-2 and SHVMD-3 systems, and the SHVMD-2 system produced the lowest specific energy consumption and the highest thermal efficiency. In addition, we tested the intermittent surface heating process, which can further improve energy performance through reducing specific electrical energy consumption in vacuum membrane distillation. This paper provides a simple and efficient membrane system for the desalination of brines.
Content may be subject to copyright.
membranes
Article
Enhanced Performance of Membrane Distillation Using Surface
Heating Process
Fei Han 1, *, Shuxun Liu 1, Kang Wang 1and Xiaoyuan Zhang 2


Citation: Han, F.; Liu, S.; Wang, K.;
Zhang, X. Enhanced Performance of
Membrane Distillation Using Surface
Heating Process. Membranes 2021,11,
866. https://doi.org/10.3390/
membranes11110866
Academic Editors: Meng Zhang,
Shujuan Meng, Xiaoyuan Zhang and
Yu Liu
Received: 18 October 2021
Accepted: 9 November 2021
Published: 11 November 2021
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2021 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
1School of Civil and Transportation Engineering, Hebei University of Technology, Tianjin 300401, China;
a1095124986@163.com (S.L.); wangkang8012@163.com (K.W.)
2Advanced Environmental Biotechnology Centre, Nanyang Environment & Water Research Institute,
Nanyang Technological University, Singapore 639798, Singapore; Zhangxiaoyuan@tju.edu.cn
*Correspondence: hanfei@hebut.edu.cn
Abstract:
Membrane distillation (MD) is a thermally driven desalination process that has excellent
application prospects in seawater desalination or hypersaline wastewater treatment, while severe
temperature polarization (TP) and the resulting relatively high energy consumption have become
principal challenges limiting the commercial application of MD. Therefore, the design of novel
systems to overcome the shortage of conventional MD requires urgent attention. Here, we developed
three surface heating vacuum membrane distillation systems, namely, SHVMD-1, SHVMD-2, and
SHVMD-3, according to the different positions of the thermal conducting layer in the cell. The
distillate flux, TP, and energy performance of these systems under different operating conditions
were investigated. All three systems showed stable performance, with a salt rejection >99.98% for
35 g/L NaCl, and the highest flux was close to 9 L/m
2·
h. The temperature polarization coefficients
were higher than unity in SHVMD-2 and SHVMD-3 systems, and the SHVMD-2 system produced
the lowest specific energy consumption and the highest thermal efficiency. In addition, we tested
the intermittent surface heating process, which can further improve energy performance through
reducing specific electrical energy consumption in vacuum membrane distillation. This paper
provides a simple and efficient membrane system for the desalination of brines.
Keywords:
membrane distillation; surface heating; hypersaline water treatment; temperature
polarization; thermal efficiency; specific energy consumption
1. Introduction
The shortage of water resources is gradually increasing worldwide [
1
3
]. The safe and
efficient utilization of saline or contaminated water is of great significance for expanding
water sources [
4
6
]. Generally, desalination technology can be divided into two categories
according to the separation method, that is, pressure-driven membrane-based methods,
represented by reverse osmosis (RO), and thermal desalination methods, represented by
multi-effect distillation (MED) and multi-stage flash (MSF) [
7
9
]. Membrane distillation
(MD) is a thermally driven liquid separation process. Water evaporates on the feed side
and is driven by the vapor pressure and temperature difference between the two sides of
the porous hydrophobic microfiltration membrane. After passing through the membrane
pores, the vapor molecules are condensed and enriched at the distillate side to achieve the
efficient separation of water and pollutants. MD has the advantages of simple heat transfer
equipment, small footprint of equipment, strong modularity and scalability [1014].
Compared with RO, MD can handle hypersaline brine as high as 180 g/L or more
at atmospheric pressure and is more resistant to fouling due to its large membrane pore
size [1518].
Conventional MD requires preheating the feed liquid to generate a vapor pressure
difference, but the reliance on the hot feed leads to the temperature of the membrane
surface being lower than that of the feed, thus resulting in temperature polarization
Membranes 2021,11, 866. https://doi.org/10.3390/membranes11110866 https://www.mdpi.com/journal/membranes
Membranes 2021,11, 866 2 of 13
(TP) [
19
21
]. TP can reduce the mass transfer driving force by 40–65%, thus reducing the
distillate flux and increasing the energy consumption [
22
,
23
]. Therefore, mitigating the
TP is of great significance for improving MD performance in terms of distillate flux and
energy consumption. Recently, a self-heated vacuum membrane distillation (VMD) system
was developed, which differs fundamentally from a conventional VMD in that heat was
directly supplied to the membrane surface. Directly providing heat to the membrane/water
interface can mitigate TP and improve the energy efficiency of the system, for example,
applying radio frequency induction to the VMD system and using induction heating to
directly heat the brine at the membrane/water interface, thus providing an effective driving
force for the MD process while reducing the effects of TP [
24
]. Coating carbon nanotubes
on the polytetrafluoroethylene (PTFE) membrane to form a conductive composite layer,
and applying a high-frequency alternating current to generate Joule heating, can make the
temperature of the membrane surface higher than the temperature of the bulk feed, which
could mitigate TP [
25
]. In addition, since MD can operate at low temperatures, solar heating
is considered an efficient and environmentally friendly method [
26
28
]. The application of
new photothermal and electrothermal materials also makes it possible to directly heat the
membrane surface [
29
]. A double-layer hydrophilic/hydrophobic membrane composed of
carbon black-loaded polyvinyl alcohol nanofibers and a polyvinylidene fluoride (PVDF)
hydrophobic membrane can absorb solar radiation to heat the cold feed [
30
]. These methods
are effective in mitigating TP, but are still limited by the velocity of and temperature
decrease in the feed [
28
33
]. In addition, the use of novel MD process designs, such as an
O-Ring VMD, could also improve the system flux [
34
]. Previous studies have demonstrated
the feasibility and effectiveness of conducting thermal energy to the membrane/water
interface [
35
37
]. However, further research such as increasing the thermal efficiency of the
system and reducing energy consumption is still needed to improve system performance.
Therefore, in this study, we optimized the configuration of thermal conducting layers
and designed three surface heating VMD systems (namely, SHVMD-1, SHVMD-2, and
SHVMD-3). We chose VMD as the basic process because there is negative pressure on the
permeate side of VMD, while TP only occurs on the feed side, which can minimize heat loss
for a stable flux [
38
]. The influence of the configuration and operating conditions of these
systems on flux are discussed. In order to further reduce the energy consumption of the
system, we conducted intermittent heating experiments by changing the heating conditions
of the external heat source [
39
,
40
]. The systems used in this paper can reverse the adverse
effects of TP, reduce energy consumption, and provide a solution for the treatment of brines
by MD.
2. Materials and Methods
2.1. Materials
A hydrophobic PTFE membrane (Shengju Tech Co, Tianjin, China) with a thickness of
50
µ
m, pore diameter of 0.2
µ
m and porosity of 81% was used in this study. In addition,
the size of the aluminum shim used as the thermal conducting layer of the system was
11 cm
×
25 cm
×
0.08 mm. We used 35 g/L NaCl solution as the feed to test the desalination
performance of these three VMD systems; sodium chloride (AR grade) was purchased
from Damao Reagent Co., Ltd., Tianjin, China.
2.2. Surface Heating System Design
The surface heating systems were designed as shown in Figure 1a. Heat was delivered
to the feed flow channel or the hydrophobic membrane surface by heat conduction and
convection through a metallic thermal conducting layer. An aluminum shim was chosen as
the thermal conducting layer due to its good thermal conductivity, excellent mechanical
strength, and low cost.
Membranes 2021,11, 866 3 of 13
Figure 1.
Schematic diagram of the experimental setup for (
a
) schematic of a laboratory-scale surface heating VMD system;
(
b
) SHVMD-1 system: the aluminum shim was embedded into the feeding channel; (
c
) SHVMD-2 system: the aluminum
shim was placed close to the membrane surface of the feed side; (
d
) SHVMD-3 system: double-layer shims, one of which
was embedded into the feeding channel, and the other was placed close to the membrane surface.
Three configurations of thermal conducting layers were studied in this paper. Connect
one end of the aluminum shim to an external heat source and embed the other end of the
shim into the feeding channel of the cell to compose the SHVMD-1 system (Figure 1b).
Connect one end of the aluminum shim with vent pores to an external heat source and
place the other end of the shim close to the membrane surface of the feed side in the cell to
constitute the SHVMD-2 system (Figure 1c). Simultaneously place the aluminum shims in
the SHVMD-1 and SHVMD-2 into the cell to assemble the SHVMD-3 system (Figure 1d).
2.3. Surface-Heating VMD Experiments
The experiments were performed using laboratory-scale VMD systems (Figure 1a),
which were composed of a heat source, feeding pump, cell, vacuum, and condenser.
The thermal power provides energy, and the thermal conducting layer plays the role of
transferring heat into the cell. A flat sheet PTFE membrane with an effective area of 40 cm
2
(4 cm
×
10 cm) was installed in the cell. The height of the feed and distillate flow channels
was 4 mm. Pieces of aluminum (Al) shim or Al shim with vent pores were used as the
thermal conducting layer. A peristaltic pump with temperature-resistant tubing circulated
the feed solution, and the flow velocity was controlled by the pump controller. A vacuum
pump generated a vacuum in the range of 0~100 kPa on the distillate side of the membrane.
As shown in Figure 1a, feed solution was circulated in the feed side by the peri-
staltic pump (Lead Fluid, BT101T, Baoding, China). The external heat source (Chang Run,
XMT615, Shenzhen, China) transferred heat to the cell through the thermal conducting
layer. The heat was delivered to the membrane/water interface for evaporation. Water
vapor diffused through the membrane pores and entered the permeate side via a vacuum
pump (Yong Hao, 2XZ-4, Linhai, China); it was then condensed by a condenser, which
was driven by another peristaltic pump (Lead Fluid, BT600-2J, Baoding, China). The tem-
perature inside the cell and flow channel was measured using thermocouples (Shenhua,
Type K, Shenzhen, China). Additionally, the conductivity of feed and produced water
was measured using a thunder magnetic conductivity meter (DDS-11A, Ray Magnetic,
Membranes 2021,11, 866 4 of 13
Shanghai, China). The experiment was repeated three times, and the results are reported
as averages.
2.4. System Performance
2.4.1. Distillate Flux
The distillate flux was calculated using Equation (1) [20]:
J=m
ρAt (1)
where J(L/m2·h) is the distillate flux; m(kg) and ρ(kg/L) are the quality reduction and
density of the feed solution, respectively; A(m
2
) is the effective membrane area; and t(h)
is the operation time.
2.4.2. Rejection
The rejection rate can be expressed in terms of the conductivity of the feed. In this
experiment, a thunder magnetic conductivity meter was used to measure the conductivity
as follows:
R=(kfkp)
kf
×100% (2)
where Ris the rejection rate (%); k
f
(mS/cm) and k
p
(mS/cm) are the conductivities of the
feed and permeate, respectively.
2.4.3. Temperature Polarization Coefficient
The temperature polarization coefficient (TPC) was used to quantify the severity of TP,
which can be defined by the ratio of the membrane surface temperature on the feed side to
the temperature of the bulk feed in the VMD system, and can be calculated as follows:
TPC =Tmf
Tbf
(3)
where T
mf
(
C) is the membrane surface temperature on the feed side; T
bf
(
C) is the bulk
temperature of the feed.
2.4.4. Heat Transfer Performance
The surface heating VMD does not need to preheat the feed solution but requires
transfer heat to the membrane/water interfaces through the thermal conducting layer. In
the heat transfer process, the external heat source conducts heat to the heat conducting
layer, then convective heat exchange occurs between the heat conducting layer and the feed.
The convective heat exchange that occurs in the vacuum can usually be ignored [4044].
Q
in
(W) is the transfer of heat from the heat source to the thermal conducting layer,
which can be calculated using the following Equation:
Qin =AmKm
T
d(4)
where A
m
is the cross-section area of the shim (m
2
); K
m
is the heat transfer coefficient of the
aluminum shim, 237 W/m
·
K; d(m) and
T(K) are the heat conduction distance on the
shim and the temperature difference of the distance, respectively.
Q
v
is the heat absorbed as latent heat when water evaporates, which can be calculated
as follows:
Qv =J AHv (5)
where Hvis the latent heat of evaporation of water, 2357.6 kJ/kg.
Membranes 2021,11, 866 5 of 13
The system thermal efficiency (TE) refers to the proportion of the total heat required
for evaporation, which is calculated as follows:
TE(%) = Qv
Qin
(6)
SEC (kWh/L) is defined as the amount of total energy supplied to produce a unit
mass of pure water, which is calculated as in Equation (9) [37,45]:
SEC =STEC +SEEC (7)
where STEC (kWh/L) and SEEC (kWh/L) are the specific thermal energy consumption
and the specific electrical energy consumption, respectively, which can be calculated as in
Equations (8) and (9) [46]:
STEC =Qin ρ
JA (8)
SEEC =Eρ
J A h(9)
where E(kJ/s) presents the rate of electrical energy input, including the vacuum pump,
peristaltic pump, and condenser.
3. Results and Discussion
3.1. Effect of Operating Conditions on Distillate Flux
Here, we discuss the effects of the operating conditions, including the heat source
temperature, crossflow velocity, vacuum, and feed salinity, on the distillate flux and the
corresponding heat input of the three systems.
Figure 2a shows that with the increase in heat source temperature from 50
C
±
1
C
to 200
C
±
1
C, the flux of the three systems increased gradually. The reason for this is
that the driving force of VMD mass transfer is mainly derived from the transmembrane
vapor pressure difference, which is composed of the saturated vapor pressure difference
and vacuum level [
21
]. Under constant vacuum conditions, the transmembrane pressure
difference is determined by the saturated vapor pressure difference, while the saturated
vapor pressure and the surface temperature of the membrane are positively correlated ac-
cording to the Antoine Equation [
43
]. Therefore, as the transmembrane pressure difference
increases with the heat source temperature, the flux eventually increases.
The flux in the SHVMD-2 system was increased by 20%
±
1.8%, 19%
±
1.9%, 20%
±
2.3%, and 21%
±
2.1% at four different heat source temperatures, respectively, compared
to that in the SHVMD-1 system. This can be attributed to the direct contact between the
thermal conducting layer and the hydrophobic membrane in the SHVMD-2 system, making
the membrane surface temperature higher than the bulk feed temperature, which results in
the elimination of TP, the reduction in heat and mass transfer resistance, and the increase
in distillate flux [31].
The input heat in all systems increased with the heat source temperature (Figure 2a).
Under the same operating conditions, the flux of the SHVMD-2 system was significantly
higher than that of the SHVMD-1 system, while the heat input was almost equal in the
two systems, indicating that system configuration—in addition to heat input—plays an
essential role. In the SHVMD-1 system, the thermal conducting layer heated the feed from
the bottom of the feed channel, causing unnecessary heat loss because part of the energy
was used in the non-evaporation process. However, in the SHVMD-2 system, the heat
was directly transferred to the membrane/water interface, which can minimize the heat
loss of the non-evaporating process. This can be attributed to the fact that evaporation is a
surface process, in which water molecules at the very thin air/water interface in the feed
side, driven by their high energy state, are transported into the vapor phase [42].
Membranes 2021,11, 866 6 of 13
Figure 2.
Heat input and distillate flux in 2 h tests with different system configurations. The flux
and heat input of SHVMD-1, SHVMD-2, and SHVMD-3 were measured as a function of (
a
) heat
source temperature, (
b
) crossflow velocity, (
c
) vacuum level, and (
d
) salinity. Regarding the operating
conditions, unless specified as the variable, the heater temperature was 200
C, crossflow velocity
was fixed at 3 cm/s, vacuum level was kept at 90 kPa, and salinity was 35 g/L.
Although doubled heat was input in the SHVMD-3 system due to the use of a
dual thermal conducting layer, the increase in flux was not considerable compared to
the others, which may be related to system thermal efficiency and be discussed in the
following section.
The distillate flux of the three systems increased first and then decreased with crossflow
velocity (Figure 2b). The increase in flux can be interpreted as the increase in fluid velocity
with the increase in volumetric flow rates; thus, the convective heat transfer coefficient
developed, and the thermal boundary layer thickness decreased. The subsequent decrease
in flux can be explained as the residence time of the feed solution becoming shorter, and the
high crossflow velocity led to a decrease in heat absorbed by the feed in the cell, resulting in
a decrease in distillate flux. In addition, high crossflow velocity may lead to an increase in
pressure on the membrane surface, increasing the risk of membrane wetting [
47
]. Therefore,
the velocity used in this study was 3 cm/s.
The distillate flux increased with the vacuum level, indicating that vacuum level had
a great influence on the transmembrane pressure difference in this study (Figure 2c). An
ideal flux performance was obtained by combining the suitable temperature and vacuum
level. The flux was only slightly changed with a vacuum level of less than 85 kPa; then,
the flux increased sharply. This is because the driving force of mass transfer of VMD
is mainly derived from the transmembrane pressure difference, which is composed of
vacuum level and saturated vapor pressure differences [
41
]. Thus, a greater mass transfer
Membranes 2021,11, 866 7 of 13
driving force was obtained via an increase in the vacuum level at a certain heat source
temperature. In addition, according to the Clausius–Clapeyron equation, the boiling point
changes non-linearly with the vacuum level [
48
]. The increase in the vacuum level can
greatly reduce the boiling point, thereby increasing the distillate flux (Figure 3).
Figure 3. Changes in flux and boiling point of solution with the vacuum level.
When the salinity increased from 10 g/L to 100 g/L, the distillate flux decreased
because the partial pressure of water vapor on the membrane surface decreases with the
salinity, thus reducing the transmembrane pressure difference. However, the flux still
reached 7 L/m
2·
h at 100 g/L salinity (Figure 2d), indicating that the systems were not
sensitive to salinity. As the salinity of the solution in the thermal boundary layer on the
surface of the hydrophobic membrane gradually increased with the evaporation process
of the feed solution, concentration polarization occurred. When the concentration of feed
near the thermal boundary layer reaches saturation, salt crystals will be produced, thus
resulting in membrane scaling and a decrease in the distillate flux.
All of the above operating conditions had little effect on the rejection rate, which
remained above 99.98%, showing the inherent advantages of MD in the treatment of brine.
3.2. TPC
The TPC decreased with heat source temperature in the SHVMD-1 system because
the thermal conducting layer conducted the heat to the feed, then transferred it to the
membrane surface, resulting in the temperature in the feed being higher than that in
the membrane surface (Figure 4). In addition, the convective heat transfer rate and flux
increased with the increase in temperature [
49
]. Therefore, the evaporation took away more
heat and reduced the temperature of the membrane surface, leading to a decrease in TPC.
Figure 4. TPC changes with heat source temperature in different systems.
Membranes 2021,11, 866 8 of 13
However, in the SHVMD-2 and SHVMD-3 systems, the thermal conducting layer was
directly in contact with the membrane surface. The heat taken away by the evaporation
was supplemented to the thermal boundary layer by the thermal conducting layer, which
overcame the negative impact of the reduction in the driving force at the surface of the
membrane due to evaporation and heat absorption of the feed. In addition, the thermal
conducting layer also played a role in heating the hydrophobic membrane. Although both
the membrane surface and the feed were heated up, the circulation of the feed made it
difficult to accumulate the heat transferred from the thermally conductive layer to the feed.
Moreover, the hydrophobic membrane was a poor conductor of heat, which caused the
membrane surface temperature to be higher than the feed temperature. Therefore, TPC
gradually increased with the heat source temperature.
In contrast to the SHVMD-1 system, the TPC values of SHVMD-2 and SHVMD-3
systems both overcame the bottleneck of unity, indicating that the direct heating of the
membrane surface by the thermal conducting layer can effectively alleviate the TP. Due to
the double thermal conducting layer in the SHVMD-3 system, the temperature difference
between the membrane surface and the feed was less than that in the SHVMD-2 system;
thus, the TPC in the SHVMD-3 system was lower than that in the SHVMD-2 system.
The above results showed that both SHVMD-2 and SHVMD-3 have excellent allevi-
ating effects on TP, and the SHVMD-2 system performed best. In addition, the TP in the
MD process was not affected by a single factor but by comprehensive manifestations under
multiple conditions. Therefore, in practical applications, it is necessary to balance water
production and energy consumption to obtain the appropriate operating temperature.
3.3. Energy Performance
Energy performance is a critical restriction for the application of MD, so it is necessary
to analyze the specific energy consumption and thermal efficiency. We tested the SEC and
TE of the three systems at different heat source temperatures.
Results showed that as the temperature of the heat source increased, the SEC in each
system increased slightly, while the TE decreased significantly. The SHVMD-3 system had
a higher SEC and the lowest TE (Figure 5). The reason for this is that the heat transferred
into the system can be divided into three parts: one is used for evaporation, another is
used to raise the temperature of the feed, and the other is lost to the environment. The
temperature of the feed rises with the heat input, leading to an increase in the three parts.
However, a high proportion of heat used for evaporation is required to increase TE. The
high heat input leads to more heat being taken away by the hot feed circulation and the
environment. This can explain why it has the highest heat input, whereas the flux was not
considerable, as discussed in Section 3.1.
Figure 5.
SEC and TE of the three systems under different heat source temperatures. All tests were
performed with a feed salinity of 35 g/L, crossflow velocity of 3 cm/s, and vacuum level of 90 kPa.
Membranes 2021,11, 866 9 of 13
Conversely, the SHVMD-2 system had the lowest SEC and the highest TE. Under the
same heat source temperature, the TE of the SHVMD-2 system was higher than that of
the SHVMD-1 system, while the heat input was almost the same because the SHVMD-2
system could minimize non-evaporative heat loss.
3.4. Intermittent Surface Heating Process
We found in our study that when the heat source was turned off for a period of
time, the flux only decreased slightly, and more importantly, the energy consumption
of the system was reduced. Therefore, we improved the above process and proposed a
new process called intermittent surface heating. For the SHVMD-2 system, we conducted
experiments at heat source temperatures of 50
C, 100
C, and 200
C, respectively. The
SHVMD-2 system was selected because it can minimize heat loss in the intermittent surface
heating process.
The procedure of the intermittent heating experiment can be divided into three stages.
In the first stage, the system was started, and the temperature of the heat source was
increased to the set temperature until stable, which required approximately 60 min. Then,
heating was stopped for 60 min and, using the residual heat to maintain the operation of
the system, the flux changes with time were recorded. During this time, the heat source
temperature decreased by 41.3%, 60.4%, and 65.8% at the initial temperature of 50
C,
100 C
, and 200
C, while the flux only decreased by 8.4%, 7.4%, and 7.0%, respectively
(Figure 6a–c). Although the temperature of the heat source decreased by more than a half,
the flux decreased slightly, indicating that the heat transfer from the heat source to the
cell can provide the driving force required for the operation of the system. This urged
us to reduce energy consumption at the cost of a small amount of flux, which required
verification by calculating SEC.
Figure 6.
Changes in heat source temperature and flux with time in the intermittent heating process
at heat source temperatures of (a) 50 C, (b) 100 C, and (c) 200 C.
Membranes 2021,11, 866 10 of 13
The SEC of the intermittent surface heating experiment increased with the temperature
of the heat source (Figure 7), which was consistent with the results of the normal surface
heating system. By subdividing SEC into SEEC and STEC, the values of SEEC decreased
with the increase in the heat source temperature. This is because the high-temperature
system has a high flux, and the electric energy consumed per unit of water is reduced when
the electric energy consumption is basically the same. However, the STEC showed the
opposite trend because the higher the heat source temperature, the more heat energy needs
to be input. Although the flux increased, the required heat input increased more, which
led to an increase in STEC.
Figure 7.
Comparison of SEC with and without intermittent heating at different heat source
temperatures.
For intermittent surface heating, when the heat source temperature was 50
C,
100 C
,
and 200
C, the SEC decreased by 8.1%, 18.0%, and 20.7%, while the STEC declined by
37.1%, 38.2%, and 40%, respectively (Figure 7). Therefore, intermittent heating had a major
influence on the STEC of the system, because the energy consumption was greatly reduced
after turning off the heat source. In addition, the higher the heat source temperature, the
better the optimization of system energy consumption by intermittent heating.
These results indicate that intermittent heating systems can reduce SEC by up to 20%
at the expense of less than 10% flux. This may lead to new breakthroughs in decreasing the
energy consumption of MD.
4. Conclusions
In this study, three surface heating vacuum membrane distillation systems were de-
veloped for the desalination of brines, and the effects of operating conditions on the system
performance were also discussed. The results showed that salt rejection in all systems was
above 99.98% at a feed salinity of 35 g/L, a crossflow velocity of 3 cm/s, and a vacuum
level of 90 kPa. Furthermore, it was found that the surface heating could significantly
reduce the TP while increasing the distillate flux. The performance of SHVMD-1 was not
as high as that of SHVMD-2 and SHVMD-3 in terms of flux and energy consumption.
The SHVMD-2 system was shown to be effective in eliminating the negative impact of
TP on MD, and achieved the highest thermal efficiency. On the other hand, a higher flux
and heat loss were observed in SHVMD-3 as compared to SHVMD-1 and SHVMD-2. In
practical applications, a trade-off between flux and energy consumption should be taken
into account in the selection of SHVMD-2 or SHVMD-3. Intermittent heating resulted in a
flux decline of less than 10% in exchange for a reduction of up to 20% SEC, indicating the
effectiveness of the intermittent heating in reducing system energy consumption. It should
also be pointed out that the efficacy of intermittent surface heating in other MD processes
with lower thermal efficiency still need further verification. It is expected that this study
can offer useful insights into energy-efficient MD processes.
Membranes 2021,11, 866 11 of 13
Author Contributions:
Conceptualization, F.H.; methodology, F.H., S.L. and K.W.; investigation,
S.L. and K.W.; data curation, S.L. and K.W.; writing—original draft preparation, S.L. and K.W.;
writing—review and editing, F.H. and X.Z.; supervision, F.H.; funding acquisition, F.H. All authors
have read and agreed to the published version of the manuscript.
Funding:
This research was funded by the Natural Science Foundation of Hebei Province, grant
number E2021202002.
Acknowledgments:
The authors thank the Natural Science Foundation of Hebei Province for funding
this study.
Conflicts of Interest: The authors declare no conflict of interest.
Nomenclature
Aeffective area of the membrane, m2
Amcross-section area of the shim, m2
Cspecific heat capacity, J/kg·C
dheat conduction distance on the shim, m
Erate of electrical energy input, kJ/s
Hvlatent heat of evaporation, kJ/kg
Jpermeate flux, L/m2·h
kfconductivity of the feed, mS/cm
kpconductivity of the permeate, mS/cm
Kmcoefficient of thermal conductivity, W/m·K
mmass reduction in feed, kg
Qin total heat transfer to the system, W
Qvheat transfer across the VMD, W
Rrejection rate, %
SEC specific energy consumption of the system, kWh/L
SEEC specific electrical energy consumption, kWh/L
STEC specific thermal energy consumption, kWh/L
toperation time, h
Ttemperature difference of the distance, K
Tbf bulk temperature of the feed, C
Tmf membrane surface temperature on feed side, C
TE thermal efficiency, %
TPC temperature polarization coefficient
ρdensity of feed solution, kg/L
References
1.
Giwa, A.; Dufour, V.; Al Marzooqi, F.; Al Kaabi, M.; Hasan, S.W. Brine management methods: Recent innovations and current
status. Desalination 2017,407, 1–23. [CrossRef]
2. Hoekstra, A.Y. Water scarcity challenges to business. Nat. Clim. Chang. 2014,4, 318–320. [CrossRef]
3.
Jones, E.; Qadir, M.; van Vliet, M.T.H.; Smakhtin, V.; Kang, S.-M. The state of desalination and brine production: A global outlook.
Sci. Total. Environ. 2019,657, 1343–1356. [CrossRef] [PubMed]
4.
Salih, H.H.; Dastgheib, S.A. Treatment of a hypersaline brine, extracted from a potential CO
2
sequestration site, and an industrial
wastewater by membrane distillation and forward osmosis. Chem. Eng. J. 2017,325, 415–423. [CrossRef]
5.
Sanmartino, J.A.; Khayet, M.; Garcia-Payo, M.C.; El-Balcouri, H.; Riaza, A. Treatment of reverse osmosis brine by direct contact
membrane distillation: Chemical pretreatment approach. Desalination 2017,420, 79–90. [CrossRef]
6.
Gude, V.G. Desalination and water reuse to address global water scarcity. Rev. Environ. Sci. Bio-Technol.
2017
,16, 591–609.
[CrossRef]
7.
Ghaffour, N.; Missimer, T.M.; Amy, G.L. Technical review and evaluation of the economics of water desalination: Current and
future challenges for better water supply sustainability. Desalination 2013,309, 197–207. [CrossRef]
8.
Deshmukh, A.; Boo, C.; Karanikola, V.; Lin, S.; Straub, A.P.; Tong, T.; Warsinger, D.M.; Elimelech, M. Membrane distillation at the
water-energy nexus: Limits, opportunities, and challenges. Energy Environ. Sci. 2018,11, 1177–1196. [CrossRef]
9.
Ali, E.; Orfi, J.; AlAnsary, H.; Soukane, S.; Elcik, H.; Alpatova, A.; Ghaffour, N. Cost analysis of multiple effect evaporation and
membrane distillation hybrid desalination system. Desalination 2021,517, 115258. [CrossRef]
10.
Mezher, T.; Fath, H.; Abbas, Z.; Khaled, A. Techno-economic assessment and environmental impacts of desalination technologies.
Desalination 2011,266, 263–273. [CrossRef]
Membranes 2021,11, 866 12 of 13
11.
Suarez, A.; Fernandez, P.; Ramon Iglesias, J.; Iglesias, E.; Riera, F.A. Cost assessment of membrane processes: A practical example
in the dairy wastewater reclamation by reverse osmosis. J. Membr. Sci. 2015,493, 389–402. [CrossRef]
12.
Chen, Q.; Muhammad, B.; Akhtar, F.H.; Ybyraiymkul, D.; Muhammad, W.S.; Li, Y.; Ng, K.C. Thermo-economic analysis and
optimization of a vacuum multi-effect membrane distillation system. Desalination 2020,483, 114413. [CrossRef]
13.
Burhan, M.; Shahzad, M.W.; Ybyraiymkul, D.; Oh, S.J.; Ghaffour, N.; Ng, K.C. Performance investigation of MEMSYS vacuum
membrane distillation system in single effect and multi-effect mode. Sustain. Energy Technol. Assess. 2019,34, 9–15. [CrossRef]
14.
Chen, Q.; Burhan, M.; Akhtar, F.H.; Ybyraiymkul, D.; Shahzad, M.W.; Li, Y.; Ng, K.C. A decentralized water/electricity
cogeneration system integrating concentrated photovoltaic/thermal collectors and vacuum multi-effect membrane distillation.
Energy 2021,230, 120852. [CrossRef]
15.
Mericq, J.-P.; Laborie, S.; Cabassud, C. Vacuum membrane distillation of seawater reverse osmosis brines. Water Res.
2010
,44,
5260–5273. [CrossRef] [PubMed]
16. Alklaibi, A.M.; Lior, N. Membrane-distillation desalination: Status and potential. Desalination 2005,171, 111–131. [CrossRef]
17.
Choi, Y.; Kim, S.-H.; Lee, S. Comparison of performance and economics of reverse osmosis, membrane distillation, and pressure
retarded osmosis hybrid systems. Desalin. Water Treat. 2017,77, 19–29. [CrossRef]
18.
Alkhudhiri, A.; Hilal, N. Air gap membrane distillation: A detailed study of high saline solution. Desalination
2017
,403, 179–186.
[CrossRef]
19.
Khayet, M. Membranes and theoretical modeling of membrane distillation: A review. Adv. Colloid Interface Sci.
2011
,164, 56–88.
[CrossRef]
20.
Drioli, E.; Ali, A.; Macedonio, F. Membrane distillation: Recent developments and perspectives. Desalination
2015
,356, 56–84.
[CrossRef]
21. Curcio, E.; Drioli, E. Membrane distillation and related operations—A review. Sep. Purif. Rev. 2005,34, 35–86. [CrossRef]
22.
Suleman, M.; Asif, M.; Jamal, S.A. Temperature and concentration polarization in membrane distillation: A technical review.
Desalin. Water Treat. 2021,229, 52–68. [CrossRef]
23. Yang, L.; Ding, Z.; Ma, R. Study on temperature polarization in membrane distillation. Membr. Sci. Technol. 2004,24, 4–9.
24.
Anvari, A.; Kekre, K.M.; Yancheshme, A.A.; Yao, Y.; Ronen, A. Membrane distillation of high salinity water by induction heated
thermally conducting membranes. J. Membr. Sci. 2019,589, 117253. [CrossRef]
25.
Dudchenko, A.V.; Chen, C.; Cardenas, A.; Rolf, J.; Jassby, D. Frequency-dependent stability of CNT Joule heaters in ionizable
media and desalination processes. Nat. Nanotechnol. 2017,12, 557–563. [CrossRef]
26.
Buenaventura Pouyfaucon, A.; Garcia-Rodriguez, L. Solar thermal-powered desalination: A viable solution for a potential market.
Desalination 2018,435, 60–69. [CrossRef]
27.
Wu, J.; Zodrow, K.R.; Szemraj, P.B.; Li, Q. Photothermal nanocomposite membranes for direct solar membrane distillation. J.
Mater. Chem. A 2017,5, 23712–23719. [CrossRef]
28.
Soomro, M.I.; Kim, W.-S. Performance and economic investigations of solar power tower plant integrated with direct contact
membrane distillation system. Energy Convers. Manag. 2018,174, 626–638. [CrossRef]
29.
Kabeel, A.E.; Abdelgaied, M.; El-Said, E.M.S. Study of a solar-driven membrane distillation system: Evaporative cooling effect on
performance enhancement. Renew. Energy 2017,106, 192–200. [CrossRef]
30.
Dongare, P.D.; Alabastri, A.; Pedersen, S.; Zodrow, K.R.; Hogan, N.J.; Neumann, O.; Wu, J.; Wang, T.; Deshmukh, A.; Elimelech,
M.; et al. Nanophotonics-enabled solar membrane distillation for off-grid water purification. Proc. Natl. Acad. Sci. USA
2017
,114,
6936–6941. [CrossRef]
31.
Anvari, A.; Yancheshme, A.A.; Kekre, K.M.; Ronen, A. State-of-the-art methods for overcoming temperature polarization in
membrane distillation process: A review. J. Membr. Sci. 2020,616, 118413. [CrossRef]
32.
Banat, F.; Jwaied, N.; Rommel, M.; Koschikowski, J.; Wieghaus, M. Performance evaluation of the “large SMADES” autonomous
desalination solar-driven membrane distillation plant in Aqaba, Jordan. Desalination 2007,217, 17–28. [CrossRef]
33.
Chen, Y.-H.; Li, Y.-W.; Chang, H. Optimal design and control of solar driven air gap membrane distillation desalination systems.
Appl. Energy 2012,100, 193–204. [CrossRef]
34.
Alanezi, A.A.; Abdallah, H.; El-Zanati, E.; Ahmad, A.; Sharif, A.O. Performance investigation of O-ring vacuum membrane
distillation module for water desalination. J. Chem. 2016,2016, 11. [CrossRef]
35.
Huang, L.; Pei, J.; Jiang, H.; Hu, X. Water desalination under one sun using graphene-based material modified PTFE membrane.
Desalination 2018,442, 1–7. [CrossRef]
36.
Cui, L.; Zhang, P.; Xiao, Y.; Liang, Y.; Liang, H.; Cheng, Z.; Qu, L. High rate production of clean water based on the combined
photo-electro-thermal effect of graphene architecture. Adv. Mater. 2018,30, 1706805. [CrossRef] [PubMed]
37.
Wang, J.; Liu, Y.; Rao, U.; Dudley, M.; Ebrahimi, N.D.; Lou, J.; Han, F.; Hoek, E.M.V.; Tilton, N.; Cath, T.Y.; et al. Conducting
thermal energy to the membrane/water interface for the enhanced desalination of hypersaline brines using membrane distillation.
J. Membr. Sci. 2021,626, 119188. [CrossRef]
38.
Lokare, O.R.; Tavakkoli, S.; Khanna, V.; Vidic, R.D. Importance of feed recirculation for the overall energy consumption in
membrane distillation systems. Desalination 2018,428, 250–254. [CrossRef]
39.
Wang, D.C.; Wu, J.Y. Influence of intermittent heat source on adsorption ice maker using waste heat. Energy Convers. Manag.
2005
,
46, 985–998. [CrossRef]
Membranes 2021,11, 866 13 of 13
40.
Wang, D.C.; Wu, J.Y.; Shan, H.G.; Wang, R.Z. Experimental study on the dynamic characteristics of adsorption heat pumps driven
by intermittent heat source at heating mode. Appl. Therm. Eng. 2005,25, 927–940.
41. Alkhudhiri, A.; Darwish, N.; Hilal, N. Membrane distillation: A comprehensive review. Desalination 2012,287, 2–18. [CrossRef]
42.
Zhang, L.; Tang, B.; Wu, J.; Li, R.; Wang, P. Hydrophobic light-to-heat conversion membranes with self-healing ability for
interfacial solar heating. Adv. Mater. 2015,27, 4889–4894. [CrossRef] [PubMed]
43.
Olatunji, S.O.; Camacho, L.M. Heat and mass transport in modeling membrane distillation configurations: A review. Front.
Energy Res. 2018,6, 130. [CrossRef]
44.
Qtaishat, M.; Matsuura, T.; Kruczek, B.; Khayet, M. Heat and mass transfer analysis in direct contact membrane distillation.
Desalination 2008,219, 272–292. [CrossRef]
45.
Miladi, R.; Frikha, N.; Kheiri, A.; Gabsi, S. Energetic performance analysis of seawater desalination with a solar membrane
distillation. Energy Convers. Manag. 2019,185, 143–154. [CrossRef]
46.
Luo, A.; Lior, N. Critical review of membrane distillation performance criteria. Desalin. Water Treat.
2016
,57, 20093–20140.
[CrossRef]
47.
Laqbaqbi, M.; Antonio Sanmartino, J.; Khayet, M.; Garcia-Payo, C.; Chaouch, M. Fouling in membrane distillation, osmotic
distillation and osmotic membrane distillation. Appl. Sci. 2017,7, 334. [CrossRef]
48.
Koutsoyiannis, D. Clausius-Clapeyron equation and saturation vapour pressure: Simple theory reconciled with practice. Eur. J.
Phys. 2012,33, 1021–1022. [CrossRef]
49.
Alsaadi, A.S.; Francis, L.; Amy, G.L.; Ghaffour, N. Experimental and theoretical analyses of temperature polarization effect in
vacuum membrane distillation. J. Membr. Sci. 2014,471, 138–148. [CrossRef]
... Han et al. [49] used aluminum shim as thermal conducting layer in VMD desalination. Three different locations of the shim inside the cell were investigated: into the feed channel (SHVMD-1), close to the membrane surface (SHVMD-2), and both into the feed channel and close to the membrane (SHVMD-3). ...
Article
Full-text available
Membrane distillation (MD) is a thermal-based membrane operation with high potential for the treatment of aqueous streams. However, its implementation is limited and only few examples of MD pilots can be found in desalination. One of the reasons behind this is that MD requires thermal energy for promoting the evaporation of water, which implies higher energy consumption with respect to pressure-driven membrane operations, like reverse osmosis (RO). Recently, among the different methods investigated to improve the thermal efficiency of MD, attempts for obtaining a localized heating of the feed, close to the membrane surface, were carried out. This work reviews experimental activities on the topic, dealing with both modified membranes, used under solar irradiation or coupled to an electric source, and specifically designed heated modules. The main results are reported and points of action for further optimization are identified. In particular, although at an early stage, this type of approach led to improvements in membrane flux and to a reduction of energy consumption with respect to conventional MD. Nevertheless, long tests to ensure a stable performance time, the optimization of operating conditions, the development of methods to control fouling issues, and the identification of the best module design, together with the scale-up of membranes/modules developed, represent the main research efforts needed for future implementation of localized heating strategy.
Article
Membrane distillation (MD) is a highly promising method for desalinating water with high salt content. However, a major challenge faced by this technology is its high energy consumption, which is caused by the need to heat up the saline feed solution. Hence, this study aimed to incorporate multi-walled carbon nanotubes (MWCNTs) into the polysulfone (PSF) matrix to fabricate composite photothermal MD membranes. These membranes have the capability to function independently without the need for an external energy source. Various quantities of MWCNTs ranging from 0.5 to 1.5% (by weight) were added to the PSF casting solution. This led to the creation of composite membranes with improved photothermal characteristics. The PSF and PSF/MWCNTs composite membranes were analyzed using scanning electron microscopy (SEM), FTIR spectroscopy, thermal gravimetric analysis, and contact angle measurements. The concentration of MWCNTs that yielded the highest photothermal efficiency was found to be 1% (by weight). The membranes were assessed using a photothermal membrane distillation (PMD) device, where the operational conditions and parameters were investigated. The presence of MWCNTs resulted in an increase in the surface temperature of the membrane to 67 °C when placed 10 cm away from a 200 W light source. In addition, the inclusion of MWCNTs led to a rise in the contact angle measurement from 80° to 112°, as well as an improvement in the liquid entry pressure (LEP) from 25.5 to 52 psi. Moreover, it significantly impacted the improvement of membrane production in the presence of light, while also exhibiting a high degree of operational efficiency.
Article
There have been tremendous advances in membrane distillation (MD) since the concept was introduced in 1961: new membrane designs and process configurations have emerged, and its commercial viability has been evaluated in several pilot-scale studies. However, its high energy consumption has hindered its commercialization. One of the most promising ways to overcome this obstacle is to develop more energy-efficient membrane modules. The MD research community has therefore developed diverse new module configurations for hollow fiber and flat sheet membranes that increase the thermal energy efficiency of MD by minimizing thermal polarization, increasing mass transfer across the membrane, and improving heat recovery from the condensed vapor. This review summarizes the progress made in the design of hollow fiber and flat sheet membrane modules for MD applications. It begins with a brief introduction to MD and its configurations before describing developments in module fabrication and highlighting key areas where further research is needed.
Article
Full-text available
Thermoelectric membrane distillation has shown promise as a new membrane distillation technique capable of improving energy consumption metrics. This study features an experimental design approach to investigating the performance of a thermoelectric membrane distillation system. Screening and full factorial designs were implemented in Minitab 16 to determine the optimal process conditions for minimizing the specific energy consumption of the system. The process parameter with the most significant impact on the specific energy consumption of thermoelectric membrane distillation systems was determined and a mathematical model for predicting the specific energy consumption was derived. The study showed that adjusting the feed flowrate, the most influential continuous parameter, from a sub-optimal level to an optimal level, while keeping other process variables at their optimal levels, could lead to a 34% reduction in the system's specific energy consumption. At the optimized process parameters of the thermoelectric membrane distillation system, the minimized specific energy consumption fell about 35% below the threshold value of 1,000 kWh/m³ found among the efficient membrane distillation systems in the literature.•Thermoelectric heat exchanger provides the driving force for the membrane distillation process •Seven process variables are assumed to influence the energy consumption of the distillation process •The variables are screened before being analyzed in a full factorial experimental design
Article
Membrane distillation (MD) is an emerging technology for the brine desalination. Directly heating feed liquid at the membrane-water interface for improving the performance of MD has been demonstrated, while the problem of membrane scaling remained. Herein, carbon nanotube-based electroactive membranes were prepared and characterized by a variety of methods. It was then used in a conductive heating vacuum membrane distillation system. We firstly increased the flux by optimizing system configuration, placing a spacer in the feed channel, and setting vent holes on the thermal conducting layer. Then, aiming at silicates and sulfates scaling in MD, the electroactive membranes and titanium were used as working electrodes and the counter electrode, respectively, to discuss the electrochemical anti-scaling performance and mechanism. Results showed that the silicate scale on the membrane surface can be effectively cleaned by using the electroactive membrane as the cathode and applying a potential of -5V to destroy the structure of the silicate through an electrochemical reaction. With the electroactive membrane as the anode, the scaling of calcium sulfate can be mitigated by providing a continuous potential through electrostatic repulsion. This paper may provide further insights into inorganic scaling control in conductive heating vacuum membrane distillation process.
Article
Membrane distillation (MD) has attracted extensive attention as a technology to solve water scarcity in the fields of desalination. Surface-heated membrane distillation techniques that have emerged in recent years have improved flux due to the elimination of temperature polarization, but still suffer from membrane fouling. The use of electrochemical techniques for in-situ cleaning of membrane fouling in MD is expected to be an effective solution. Herein, electroactive membranes were prepared by loading mixtures of reduced graphene oxide (rGO) and carbon nanotube (CNT) with four mass ratios on polytetrafluoroethylene (PTFE) hydrophobic membranes. A series of analytical methods, such as scanning electron microscopy, atomic force microscopy, pore size analysis, water contact angle, Raman spectroscopy, membrane surface zeta potential, and cyclic voltammetry curve, were employed to characterize the properties of the rGO-CNT membranes. The best separation performance and conductivity were achieved at a 2:1 mass ratio of rGO to CNTs (R2 membrane). Then, the rGO-CNT membranes were applied to a conductive heating vacuum membrane distillation system to treat humic acid (HA). Compared with the PTFE membrane, the four rGO-CNT membranes showed a significant improvement in the rejection of HA. Using the rGO-CNT membranes as the cathode and the titanium as the anode, after a certain potential was applied, the flux decreases slowly. The fluxes of all rGO-CNT membranes were maintained over 9.5 kg/(m²·h) for 10 hours of continuous operation. An increase in potential helps to increase change by alleviating membrane fouling. The highest flux and HA rejection reached 13 kg/(m²·h) and 99.8%, respectively, with R2 membrane and -5 V potential applied. The mechanism of antifouling included electrostatic repulsion and electrocatalytic oxidation.
Article
This work addresses the feasibility of a previously proposed hybrid desalination system comprising multiple effect evaporation (MEE) and membrane distillation (MD) processes. The feasibility study introduced here focuses on the impact of the hybridization of two separate processes on desalination plant economics. The water cost of the hybrid system is found to be 2.05 $/m³, which compares well with previously reported values and is 17% lower than that of standalone MEE system. The sensitivity of the water production cost to the MD feed flow rate indicated the existence of an optimal cost at a single-pass feed flow rate of 900 L/h. In addition, the sensitivity of the water cost with respect to the MD unit cost indicated a 0.0084 $/m³ change in water production cost for every unit change in the MD module cost. A modified structure of the hybrid system that leverages the thermal energy of all MEE effluents has also been suggested. With this system, the overall water recovery ratio can reach as high as 52% and the water cost can further be reduced to 1.84 $/m³.
Article
Cogeneration of electricity and freshwater by integrating photovoltaic/thermal collectors and desalination systems is one of the most promising methods to tackle the challenges of water and energy shortage in remote areas. This study investigates a decentralized water/electricity cogeneration system combining concentrated photovoltaic/thermal collectors and a vacuum multi-effect membrane distillation system. The merits of such a configuration include high compactness and improved thermodynamic efficiency. To evaluate the long-term production potential of the proposed system, a thermodynamic analysis is firstly conducted. Under the climatic conditions of Makkah, Saudi Arabia, the system can convert ∼70% of the solar irradiance into useful energy. The annual productivity of electricity and distilled water are 562 kWh and 5.25 m³, respectively, per m² of the solar collector area. Electricity and water production rates are found to be impacted by hot water flowrate, feed seawater flowrate and heat storage tank dimension, while the overall exergy efficiency stabilizes at 25-27%. Based on the production rates, an economic analysis is conducted through life-cycle cost analysis. The final desalination cost is calculated to be $0.7-4.3/m³, depending on the solar collector cost and the electricity price. The derived results will enable a more in-depth understanding of the proposed solar-driven water/electricity cogeneration system.
Article
Membrane distillation (MD) is a membrane-based thermal desalination process capable of treating hypersaline brines. Standard MD systems rely on preheating the feed to drive the desalination process. However, relying on the feed to carry thermal energy is limited by a decline of the thermal driving force as the water moves across the membrane, and temperature polarization. In contrast, supplying heat directly into the feed channel, either through the membrane or other channel surfaces, has the potential of minimizing temperature polarization, increasing single-pass water recoveries, and decreasing the number of heat exchangers in the system. When solar thermal energy can be utilized, particularly if the solar heat is optimally delivered to enhance water evaporation and process performance, MD processes can potentially be improved in terms of energy efficiency, environmental sustainability, or operating costs. Here we describe an MD process using layered composite membranes that include a high-thermal-conductivity layer for supplying heat directly to the membrane-water interface and the flow channel. The MD system showed stable performance with water flux up to 9 L/m²/hr, and salt rejection >99.9% over hours of desalinating hypersaline feed (100 g/L NaCl). In addition to bench-scale system, we developed a computational fluid dynamics model that successfully described the transport phenomena in the system.
Article
Membrane distillation (MD) technology has been mainly used as a treatment method for saline or contaminated wastewater. Despite the rapid progress in material engineering and design of novel MD systems, principal challenges as temperature polarization (TP) and high-energy consumption per unit of produced water still restrict its commercialization. Recently, TP mitigation has been addressed by modification and configuration of MD systems or by using advanced materials for membrane fabrication. These include coating thermally conductive or photonic nanomaterials on the membrane's surface or using thermally conductive metallic based membranes that impact heat dispersion along the membrane. In addition, frame-like turbulence promoters and modified feed channels were shown to lower TP by enhancing the characteristics of the feed flow. Finally, systems able to directly heat the membrane's surface without preheating the feed solution, including solar, Joule, and induction heating, were shown effective in eliminating TP due to the higher temperature at the membrane-water interface in comparison to the temperature of the bulk feed solution. The review aims to summarize recent advances made in TP mitigation for MD systems and assess their influence on distillation efficiency. We include a brief description of the TP phenomenon and its observed effects on MD and describe advanced MD processes from the aspects of low or negligible TP, high distillate flux, and improved energy efficiency.
Article
Vacuum multi-effect membrane distillation is an advanced system that possesses the features and merits of vacuum membrane distillation and multi-effect distillation. It has low operating pressure and temperature, high levels of non-volatile rejection and high energy efficiency. This study presents a thermo-economic analysis and optimization of this novel system. A thermodynamic analysis is firstly conducted to evaluate the productivity and the energy consumption under varying design and operational conditions. Special emphases are placed on the impacts of the system configuration, including the number of effects and the overall membrane area, which are rarely covered in the literature. Results reveal that there is a trade-off between the production rate and the energy consumption with respect to most of the operating parameters, e.g. the feed flowrate and the cooling water flowrate. An increase in the number of effects and the membrane area will reduce the energy consumption, but the specific permeate flux for the unit membrane area also becomes lower. To obtain the optimal parameters that minimize the desalination cost, an economic study is then carried out considering a wide range of thermal energy prices. It is observed that a higher feed flowrate, more numbers of effects and larger membrane areas are preferable when the energy price is higher. However, when thermal energy with low prices is available, lower feed flowrates and smaller membrane areas are recommended. The derived results will provide useful information on the vacuum multi-effect membrane distillation system for its future design and operation.
Article
With increase in fresh water demand and lack of fresh water resources, the current water scarcity can only be solved with seawater desalination. However, due to high dependence of current desalination technologies on fossil fuels, especially in GCC countries where the share of thermal desalination systems is dominating, the environmental sustainability is at risk. Despite high operational and maintenance cost, electricity operated membrane based reverse osmosis (RO) system provides simple configuration with less capital cost. Therefore, for future sustainable desalination, more innovative and energy efficient methods have to be sought out which will not only have the low operational cost of the thermal desalination systems but they can also have simple design and fabrication cost of membrane based systems. Vacuum membrane distillation (VMD) is a thermal distillation technique that works on the vapor pressure across the hydrophobic membrane. With the introduction of heat recovery scheme within the VMD modules in form of the multi-effect VMD operation, a detailed performance analysis of the VMD system is presented in this study under different operating conditions. The performance of system is investigated on components level with comparison between single effect and multi-effect operation.
Article
We present a novel approach for vacuum membrane distillation (VMD) based on a ‘self-heating’ membrane by a high-frequency magnetic field. The system proved efficient for desalinating a range of saline solutions while overcoming temperature polarization and heat loss associated with low permeate flux in MD processes. The system utilized induction heating (IH) for fast and contactless direct heating of the membrane surface without the need for preheating the bulk feed solution. A composite membrane with a dual hydrophilic-hydrophobic layer was fabricated by spray coating iron oxide-carbon nanotubes on a hydrophobic polytetrafluoroethylene commercial membrane. We evaluated the impact of operational conditions on the permeate flux and rejection while treating high salinity feeds (35–100 g/L NaCl). Following optimization, high permeate flux and 99% salt rejection were measured at a low inlet flow velocity (2.33 cm/min) and low vacuum (20 kPa) conditions. In addition, the specific heating energy of the system was determined to be significantly lower in comparison to the conventional VMD system under similar conditions. Interestingly, the treatment of very high-salinity solutions was shown to be efficient as a result of the IH mechanism. Temperature profiles and transport mechanisms were assessed using numerical simulation which was validated by the experimental results.
Article
In this paper, a study of the energy performance of a solar powered vacuum membrane distillation was investigated on four typical days that represent the four seasons of the year namely: the equinoxes (March 21st and September 21st) and the solstices (June 21st and December 21st), through the widely used energy criteria such as specific energy consumption, gained output ratio and heat recovery factor. The results indicate that the energy performance of the process is strongly linked to the solar irradiation and as long as the VMD is coupled with solar energy, the process could be competitive with industrialized desalination technologies such as RO, MSF and MED. Besides, the effect of the applied vacuum level on the specific energy consumption was examined. Obviously, the results reveal that for minimum SEEC an optimal vacuum pressure is between 5000 and 10000 Pa. As well a technique to optimize the energy performance by recycling hot cooling water, used for steam condensation, was discussed. An amelioration of about 35%, 26% and 31%, during the hottest day (June 21st) and of about 22%, 19% and 21%, during the coldest (December 21st) in the GOR, the SEC and the daily production could be achieved.
Article
Rising water demands and diminishing water supplies are exacerbating water scarcity in most world regions. Conventional approaches relying on rainfall and river runoff in water scarce areas are no longer sufficient to meet human demands. Unconventional water resources, such as desalinated water, are expected to play a key role in narrowing the water demand-supply gap. Our synthesis of desalination data suggests that there are 15,906 operational desalination plants producing around 95 million m³/day of desalinated water for human use, of which 48% is produced in the Middle East and North Africa region. A major challenge associated with desalination technologies is the production of a typically hypersaline concentrate (termed ‘brine’) discharge that requires disposal, which is both costly and associated with negative environmental impacts. Our estimates reveal brine production to be around 142 million m³/day, approximately 50% greater than previous quantifications. Brine production in Saudi Arabia, UAE, Kuwait and Qatar accounts for 55% of the total global share. Improved brine management strategies are required to limit the negative environmental impacts and reduce the economic cost of disposal, thereby stimulating further developments in desalination facilities to safeguard water supplies for current and future generations.