Volumetria de Precipitacion
Volumetria de Precipitacion
Volumetria de Precipitacion
PII: S0009-2509(18)30487-1
DOI: https://doi.org/10.1016/j.ces.2018.07.019
Reference: CES 14373
Please cite this article as: L-Z. Zhang, Q-W. Su, Performance manipulations of a composite membrane of low thermal
conductivity for seawater desalination, Chemical Engineering Science (2018), doi: https://doi.org/10.1016/j.ces.
2018.07.019
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Performance manipulations of a composite membrane of low thermal
1. Key Laboratory of Enhanced Heat Transfer and Energy Conservation of Education Ministry, School of
Chemistry and Chemical Engineering, South China University of Technology, Guangzhou 510640, China.
2. State Key Laboratory of Subtropical Building Science, South China University of Technology, Guangzhou
510640, China.
Abstract
Membranes of low thermal conductivity (thermal insulation) but high moisture permeability are highly
porous membranes with thermal insulation could decrease the heat losses in humidification, thus the system
energy performance can be improved. To satisfy this purpose, in this study, a PVDF/SiO 2 (Si aerogel)
composite bilayer membrane of low heat conductivity is prepared and analyzed. To solve the problem of
cracking in forming SiO2 layer, the casting solution (TEOS, H2O) is doped with a drying control chemical
additive (DCCA) N, N-dimethylformamide (DMF) to decrease the surface tensions in the gel. Then
performance analysis of the developed membranes are conducted and the membranes are characterized. The
results found that the heat conductivity is decreased from 0.08199 Wm-1K-1 to 0.04487 Wm-1K-1, with a
penalty of effective moisture diffusivity sacrificed from 4.93×10-6 m2/s to 2.42×10-6 m2/s. They are superior to
the available SiO2 blend insulation membranes. Further, the membranes can be optimized by facile structural
manipulations such as adjusting the SiO2 layer thickness to realize the desired performance while balancing
the paradox between the permeability and heat conductivity. This is the first time that heat conductivity of
Keywords: Seawater desalination; Thermal insulation; PVDF/SiO2 bilayer membrane; Low thermal
conductivity
*
Corresponding author, Tel/fax: 0086-20-87114268; Email: Lzzhang@scut.edu.cn
1
Nomenclature
Nu Nusselt number
Re Reunolds number
Sh Sherwood number
Greek letters
δ thickness (m)
ε porosity
η viscosity (Pa·s)
ρ density (kg/m3)
τ pore tortuosity
Subscripts
2
a air
i inlet
lm logarithmic
m mean, membrane
o outlet, ordinary
p pressure
t total
v vapor
1. Introduction
Sometimes we call our planet “Ocean” instead of “Earth” because over 70% of our earth’s surface is
covered by water. It seems that water is abundant, but the demand for the amount of fresh water available is
urgent (Yamali and Solmus, 2007). The majority of water on earth is seawater, leaving only 2.5% as fresh
water. Moreover, most of this fresh water exists in frozen areas such as Antarctica and Greenland. And there is
only about 1% fresh water accessible to direct human uses (Service, 2006). With the world’s population
growing rapidly, the quantity of water for humanity and other living beings is also increasing (Elimelech and
Phillip, 2011).
How can the growing demand for fresh water be met? Generally, desalination can be a solution to
overcome fresh water shortage. There are several technologies to achieve desalination (Banat et al., 2007;
El-Dessouky et al., 2004; Geng et al., 2014; Hassan et al., 1998; Matz and Zimerman, 1985; Mohamed et al.,
2006; Ortiz et al., 2006; Qtaishat and Banat, 2013): (1) Multi Stage Flash and Multi-Effect Distillation (MED).
(2) Reverse Osmosis (RO), Nano Filtration (NF) and Electro Dialysis (ED). (3) Membrane Distillation.
Among these technologies for seawater desalination, the recently announced membrane-based
below 80 °C and atmospheric pressure. Therefore, it can be driven by low-grade energy such as industrial
waste heat and solar energy (Chen and Ho, 2010; Guillén-Burrieza et al., 2011; Zhang and Li, 2017). Based on
a humidification-dehumidification process, fresh water from seawater can be obtained (Li and Zhang, 2016).
However, a lot of thermal energy is lost during the membrane humidification processes, resulting in a low
energy COP of around 0.4 for MHDD (Zhang and Li, 2017). Then decreasing the energy losses in membrane
3
humidification processes becomes the key issue for performance improvement. The lost energy mostly comes
from the vapor through porous membrane by heat conduction, so membranes of low heat conductivity are
desired. Therefore, the issue of membranes with thermal insulation has come under the scientific spotlight (Tu
et al., 2010).
Generally, there are several characteristics for membranes to be suitable for MHDD process: (1) The
membranes must be hydrophobic (non-wetting) and only vapor exists in the pores. The membrane pore size
used in MHDD is between 5 nm and 1μm (El-Bourawi et al., 2006). (2) The membranes used in the MHDD
should have high mass transfer coefficient but low thermal conductivity to reduce heat losses as much as
possible. (3) The membranes should have good anti-fouling properties to prevent being blocked and reduced
Membranes of thermal insulation are also required in other membrane distillation operations. There are
varieties of membrane distillation. They are: (1) Direct Contact Membrane Distillation (DCMD). DCMD is
widely used in desalination processes (Alves and Coelhoso, 2006; Calabro et al., 1994; Godino et al., 1997;
Hsu et al., 2002; Tomaszewska et al., 1995). But its main disadvantages are the large heat losses by
conduction. (2) Air Gap Membrane Distillation (AGMD). AGMD is efficient to reduce heat losses in
desalination (Banat and Simandl, 1998). However, the mass transfer resistant is increased by static air layer. (3)
Sweeping Gas Membrane Distillation (SGMD). SGMD is often used to remove volatile organic compounds
from solution. But it is high energy-consuming to collect the distillate (Banat and Simandl, 1998). (4) Vacuum
membrane Distillation (VMD). VMD is used to separate aqueous volatile solutions. Although the conduction
heat loss can be neglected, it is only suitable for low-concentration solutions due to its complexity (Wirth and
Cabassud, 2002; Yun et al., 2006). Generally, thicker membranes may be an option to decrease heat
conduction. However, there are conflicts between low mass transfer resistance for thinner membranes and
high heat resistance for thicker membranes (Qtaishat and Banat, 2013). Therefore, membranes of low thermal
Generally, there has existed one method to develop porous membranes of low thermal conductivity. They
are blend membranes doped with insulation materials. Chen et al. (Chen et al., 2014) fabricated membranes
from three different molecular weights PVDF via the phase inversion process to reduce thermal conductivity.
A porous blend membrane was prepared by Wang and Chung (Wang and Chung, 2015) using hydrophobic
clay particles as addititives. Efome et al. (Efome et al., 2015) fabricated hydrophobic PVDF-nanoparticles
4
membranes via adding superhydrophobic SiO2 nanoparticles into the dope. According to these methods, only
some parts of the matrixes were replaced by materials of low heat conductivity, while the structures of such
membranes were not changed. The mean free path of dry air (25 °C, 1.013×105 Pa) is 70 nm (Jennings, 1988),
so it is effective to decrease heat conduction by limiting the molecular oscillations if the pore sizes are less
than 70 nm. Regretfully the pore sizes of most of these membranes are above 100 nm, so the insulation effects
So far, researches on the preparation of silica aerogel porous films for heat insulation have been rarely
reported. Considering this, this study adopts a dip-coating procedure to fabricate a composite membrane rather
than a blend membrane. First, a porous PVDF membrane is developed as the support layer. Then an insulation
layer made of silica aerogel film is directly fabricated on the support layer. The PVDF support layer is
mechanically strong but highly conductive. The silica aerogel layer is relatively fragile but it has good
thermally insulation properties. So the combination of the two layers would develop a new membrane of low
heat conductivity. The difficulty to make aerogel films lies in its easy cracking during ambient drying
processes. The first sol-gel coating procedure to prepare meso-porous silica films was reported by Brinker and
co-workers (Brinker and Scherer, 1990). However, the obtained films were mostly granular with
two-dimensional hexagonal structures. When the substrate layer is porous, intermixing of the two layers
would occur, leading to poor mass transfer. To solve this problem, Stucky and co-workers (Zhao et al., 1998)
used low-cost commercially available PEO-PPO-PEO triblock copolymers mixing with poly-(ethylene oxide)
non-ionic surfactants via dip-coat processing to fabricate continuous channel or ordered cage arrays films.
Here we introduce a drying control chemical additive (DCCA) N, N-dimethylformamide (DMF) into the sol
systems (TEOS, H2O) to solve the problem of cracking. Although several studies (Darmawan et al., 2016;
Lenza et al., 2015) have dealt with the influences of adding DMF on the structure of SiO2 films of micro pores,
few works investigate the effects on aerogel films which have nano pores. In this paper, the effects of DMF on
gel time, the morphology and the structures of the composite aerogel films are discussed. These films with the
addition of DMF have a less-defined pore structure than those prepared in the literature. They have a
three-dimensional network structure and a wider pore size distribution between 5 nm and 50 nm. This kind of
structure permits high property of thermal-insulation. They also permit the adjustment of pore sizes to
manipulate performances. When applied to seawater desalination, the thickness and structures of different
layers can be manipulated to balance the paradox between permeability and heat conductivity.
5
2. Experimental
2.1. Materials
Tetraethoxysilica (≥98%, TEOS), ethanol (≥99.7%, EtOH), hydrochloric acid (34.46wt%, HCl),
ammonia hydroxide (25wt%, NH4OH), dimethylformamide (DMF, reagent grade), trimethyl chlorosilane
(TMCS, reagent grade) and acetyl acetone (AcAc, reagent grade) were obtained from Guangzhou Qianhui Co.,
China. Poly (vinylidene fluoride) (PVDF, Mw=700000) and polyethylene glycol 2000 (PEG2000) were
purchased from Aladdin, Shanghai, China. SiO2 aerogel particles (pore size 5~40 nm, porosity ≥90%, grain
size 40~80 nm, hydrophobic) were bought from Xiamen Nameite New Materials Technology Co., Ltd, China.
In this study PVDF hydrophobic membranes were prepared by a direct wet phase inversion method.
PVDF and PEG2000 were added into DMF with a specified mass ratio (PVDF/DMF/PEG2000, 9: 88.2: 2.8),
and the mixture was stirred at 55 °C for 8 h on a heating magnetic stirrer until fully dissolved. After the
solution was cooled down, the membrane preparation was the same as our previous publication (Zhang et al.,
2011). The prepared PVDF membrane was dried for later use.
PVDF-SiO2 hybrid membranes were made by the following method. 2 g hydrophobic SiO2 aerogel
powder and 2.8 g PEG2000 were mixed in DMF and dispersed ultrasonically for 2 h. Subsequently, PVDF
was added to form the casting solution. After that, the membrane preparation was the same as the section
2.2.1.
The PVDF/SiO2 composite bilayer membrane was prepared by two steps. Firstly, the fabrication of the
supporting PVDF layer was the same as that described in the section 2.2.1. Secondly, silica aerogel films were
6
Dip-coating steps were shown in Fig. 1. They were: (1) the fabrication of silica sols (SS) by sol-gel
methods. TEOS, EtOH, DMF, deionized water and HCl were mixed (1: 5: 0.8: 1: 1×10-5, molar ratio) and
magnetically stirred for 60 min. Secondly, EtOH, H2O and NH4OH were added into the sol solution and
magnetically stirred for 30 min. The final molar ratio of TEOS: EtOH: H2O: DMF: HCl: NH4OH was 1: 7: 2:
0.8: 10-5: 3.57×10-5. (2) Preparation of SS with appropriate viscosity. When the viscosity of the silica sols
reached 10×10-3 Pa·s, EtOH was added with a volume ratio of SS: EtOH=1: 3. Then acetyl acetone (AcAc)
was added to maintain a constant viscosity of 5~7×10-3 Pa·s. Table 1 showed the effects of different viscosity
and dip-coating speed on film quality. (3) Fabrication of silica aerogel films. The PVDF layer was immersed
in SS at a dip-coating rate of 10 cm/min for different cycles. To ensure films only on one side, SS on the back
side of PVDF were wiped off. Then they were quickly dropped into saturated EtOH solutions. In order to
prevent abrupt solvent evaporation, the coated wet gels were first immersed in EtOH: H2O (7: 3, vol.) with a
moderate NH4OH for 24 h at 60 °C and then in EtOH: TEOS (7: 3, vol.) for 16 h at 60 °C to strengthen the gel
network. Continuously, the water and ethanol were exchanged with n-hexane repeatedly until solvents were
replaced completely. After surface modification for 8 h at 50 °C in TMCS: n-hexane (1: 9, vol.), the wet gel
films were washed in n-hexane repeatedly. Finally, the wet films were dried by slow heating at 60 °C, 80 °C,
120 °C, 180 °C for 2 h respectively. The resultant membranes were PVDF/SiO2 composite bilayer
membranes.
Table 1
5~7×10-3 5 breakage
5~7×10-3 20 breakage
>7×10-3 5 breakage
>7×10-3 10 breakage
>7×10-3 20 breakage
7
Mixing of TEOS and EtOH using DMF as DCCA
Here, dip-coating rate referred to the speed of pulling out PVDF substrates from SS. Neither too fast nor
too slow dip-coating rates could lead to uniform films. As a result, to fabricate optimal silica films, the falling
speed was set as 8 cm/min and the pulling rate (dip-coating rate) was 10 cm/min. A constant viscosity of
The PVDF membrane thickness was measured by an electronic digital indicator (0~25 mm, 0.01 mm,
Beijing Huafeng Science and Technology Co., Ltd, China). Each membrane was measured five times and the
Thicknesses of SiO2 aerogel films were given by the literature (Brinker et al., 1992):
η (1)
ρ
Where δ was film thickness (m), c was constant, η was sol viscosity (Pa·s), and U was dip-coating rate
(m/s). According to the equation, thickness of aerogel films was 100 nm at a dip-coating rate of 10 cm/min.
8
Under these conditions, thicknesses of silica thin films with multi-coating cycles were calculated. For
convenience, the different membranes were marked as M1, M2, M3, M4, M5, M6 and M7, respectively. They
were listed in Table 2. M5 was validated by SEM observations of the membrane cross sections. It was
Table 2
power
M3 PVDF, TEOS 152 μm (1) PVDF layer by direct wet phase inversion
M4 PVDF, TEOS 154 μm (1) (2) the same as above; (3) 40 dip-coating cycles
M5 PVDF, TEOS 156 μm (1) (2) the same as above; (3) 60 dip-coating cycles
M6 PVDF, TEOS 158 μm (1) (2) the same as above; (3) 80 dip-coating cycles
M7 PVDF, TEOS 160 μm (1) (2) the same as above; (3) 100 dip-coating cycles
The contact angle of these membranes was measured by a JC2000Cl contact angle system (Shanghai,
China) with a 4 μL water droplet. Each sample was measured three times and the average contact angle was
used. The surface topology of membrane was characterized by Atomic force microscopy (AFM, XE-100, Park,
Korea). The scan size was 5 μm×5 μm. Scanning electron microscopy (SEM, Merlin, LEO1530VP, Germany;
JEM-2100F, Japan) was used to observe the membrane morphologies. A polycrystall X-raydiffraction (XRD,
Bruker D8 advance, Germany) was performed to analyze the membrane crystalline microstructure. The θ
angular range was selected between 10° and 90°. Infrared spectroscopies of the membranes were obtained in
KBr pellets using a Bruker Hyperion FT-IR spectrometer. The parameters of SiO2 aerogel layer were
characterized by using a Micromeritics ASAP 2020 automated system. The parameters of PVDF layer were
9
characterized by using a mercury porosimeter (Micrometrics Autopore 9500) based on the Carman-Kozeny
theory (Kaviany, 1991). The thermal conductivities of membranes were measured by the thin film module of
the Hot Disk Thermal Constants Analyzer (hot disk, TPS 2500S, Sweden). The TPS 2500S system was based
on a specially designed Wheatstone bridge with the Hot Disk sensor in one of the arms.
The membranes were designed for seawater desalination. Air humidification were realized with these
porous membranes. The whole set of device had been set up and used for research in our previous paper
(Zhang et al., 2011). The mass exchange rates through membranes (isothermal moisture exchange between
two air flows) were measured by the apparatus. The schematic diagram is depicted in Fig. 2. The width and
length of the flow channel were 2.5 cm and 7 cm, respectively. Channel pitch was 2 mm. Therefore, the
exchange area through membrane was 17.5 cm2. As diagrammed in Fig. 3, stream 1 and 2 represented humid
air and dry air, respectively. Absolute moisture content difference between the two streams was about 8.14
g/kg, and the vapor pressure difference was 1264 Pa. Five flow rates from 0.05 m3 to 0.26 m3 were tested for
each membrane. In addition, the differences between the sum of moisture content to and from the exchanger
were checked to be within ±2%. The uncertainties were calculated in our previous paper (Zhang, 2006).
Membrane
Stream 2 out Stream 2 in
10
Valve
Temperature and
By pass Humidity Sensor
Dehumidifier
Stream 2 in Stream 1 out
Pump
Flow Meter
and Controller
The total resistance in the test apparatus includes the membrane part itself and the convective resistance
on both sides of the membrane. The membrane part will not han e whether it’s in a liquid-gas system or in a
gas-gas system, as long as the correct boundary layer resistance is used. In the test, the gas –membrane
convective resistance is clarified and subtracted from the total resistance to obtain the membrane side
resistance. In future liquid-membrane-gas systems, like desalination, the overall resistance can be easily
calculated by the correct liquid side resistance, the previously obtained membrane resistance and the gas side
resistance.
The duct is a rectangular duct (2.57cm). The pitch is 2mm. When the duct length L>> Lh (the entry
length), the flow can be considered as fully developed. But in a short length channel, entrance effect should be
considered. For adequacy, care must be taken to ensure that fully developed flow prevails over most of the
channel. In the present design, the boundary layers on both sides in laminar flow will merge at a point
downstream. In rectangular ducts, the entrance length is given by (Muzychka and Yovanovich, 2009):
h . h (2)
Table 3
Changes of entrance length with Reynolds number.
V (m3/s) Re Lh (cm) Lh/L
0.05 68.06 0.28 3.96%
0.1 136.13 0.55 7.91%
0.16 217.81 0.89 12.66%
0.2 272.26 1.11 15.83%
0.26 353.94 1.44 20.58%
11
From table 3, the range of Re varies from 68.04 to 353.94, while the entry length is from 0.28 cm to 1.44
cm. When the flow velocity is below 0.26 m3/h, the ratio of entry length to whole channel is below 21%.
Under flow rates of 0.05m3/s, the ratio is below 5%. Based on this situation, it is reasonable to neglect the
entry effect and calculate the mass transfer coefficient as fully developed flow.
In this study, the duct length is 7 cm, so it can be regarded as fully developed flow. For fully developed
performed. The two flows have the same temperature at inlets, so the influences from temperature can be
neglected. Further, the two flows are in a counter flow arrangement, under such conditions, a constant mass
When the silica aerogel film is prepared by ambient drying, the most common problem is cracking. The
crack arises from the unbalanced capillary forces in pores, which are caused by the contraction of the network
structure of gels during drying. Several measures are taken to prevent cracking. By modifying the –OH groups
into non-active groups –CH3 with TMCS (Deshpande et al., 1992), the shrinkage of the gel network and
cracking are hindered effectively. Simultaneously, Using DMF as drying control chemical additives (DCCA)
is another method to further prevent cracking. In order to obtain appropriate molar ratios of DMF to TEOS,
the effects of DMF on the gel time and viscosity of the silicon sol are measured in Fig. 4.
12
20 30
18 28
16 26
14 24
Fig. 4. Effects of molar ratios of DMF to TEOS on the gelation time and viscosity of silicon sol.
Fig. 4 shows the effects of molar ratio of DMF to TEOS on the gelation time and viscosity of silicon sol.
With an increase of DMF, the gelation time is obviously extended. It is resulted from hydrogen bonding
interaction with H+ in the sol. Combined with the hydrolyzed intermediate to reduce the H+ content absorbed
on the sol particles, the electric potential (ζ) of the micelles is increased. So the barrier for the sol particles to
agglomerate would increase (Adachi and Sakka, 1988). Hence, the gel reaction slows down and the stability
of the sol is improved. Simultaneously, after 20 h of aging, it is observed that the larger the molar ratios of
DMF to TEOS are, the smaller the sol viscosity is. As a polar solvent, it is easy for DMF to have an affinity to
EtOH. Therefore, strong hydration between ethanol and water is reduced and the sol system contains more
free water. Consequently, the condensation polymerization of the sol system is limited and the sol viscosity is
decreased. Too high or too low viscosity is bad for the coating. In order to fabricate a uniform film, the
content of DMF should be controlled within the appropriate range. When nDMF/nTEOS is up to 0.8, the gelation
time basically remains stable. Therefore, the molar ratio of DMF to TEOS is chosen as 0.8 in this study.
13
(a) (b)
200 nm 200 nm
Fig. 5. The films before and after doping DMF. (a) before; (b) after.
To discuss the effects of DMF on the structure of the aerogel film, the films supported on Si wafers
before and after doping DMF were prepared. Fig. 5 shows the surface morphology of the films before and
after doping DMF. From Fig. 5 (a), the silica particles are stacked irregularly and the film surface is coarse.
After adding DMF, films of uniform pore sizes and high porosity are formed in Fig. 5 (b). This shows that
DMF can prevent gel breakage effectively and inhibit the formation of particle clusters. This is due to the
effects of DMF on silica colloidal particles. The mechanism is depicted in Fig. 6. In the gel process, when
combined with the hydrolyzed intermediate, the activities of the intermediate Si(OH)xOR4-x and the product
SixOy(OH)z are suppressed. The active participation of DMF in steric shielding around Si causes a decrease in
the colloid particle sizes. Therefore, it is effective on the promotion of polymerization and the inhibition of
coalescence. Meanwhile, because the growth of sol particles is inhibited, the pore diameter and pore volume
increase and the gel shrinkage during ambient drying decreases. Consequently aerogel films with a uniform
14
The surface contact angles (SCA) were measured to describe the hydrophobicity of the modified PVDF
Table 4
Samples M1 M2 M3 M4 M5 M6 M7
The relationship between the contact angle of membranes and wettability is given by Young's equation
(Zisman, 1964). Wetting of the surface is very favorable (hydrophilic) when a contact angle is less than 90°;
wetting of the surface is unfavorable when a contact angle is greater than 90°.
For the PVDF membranes (M1), the formed macro-porous membrane is hydrophilic with a contact angle
of 76.2° due to -OH groups of PEG2000. However, the PVDF-SiO2 hybrid blend membrane (M2) becomes
hydrophobic with the addition of hydrophobic SiO2 nanoparticles. For the PVDF/SiO2 composite bilayer
membranes (M3-M7), it is observed that the SCA changes with different number of dip-coating cycles. For
M3, because the dipped silica sols are re-dissolved in EtOH, silica films are not formed. So M3 is hydrophilic.
Membranes M4~M6 are hydrophobic. The reason lies in the methylation reaction with TMCS. The more the
solution is coated on the surface, the more complete the silica films are formed. Therefore, M5 has the largest
contact angle of 115.7°, which helps to prevent the penetration of aqueous solution on both sides of the
membrane. After dip-coated 80 cycles, more coated solution would have adverse effects. For M7, because the
coating process is too long, during which a rapid solvent evaporation makes the film cracked. So it becomes
hydrophilic again.
AFM imaging is performed to assess the membrane uniformity on top surfaces. The scan size was 5
μm×5 μm. The mean roughness (Rq) is used to indicate the roughness of the membrane surfaces, which is
listed in Table 5.
15
Table 5
Samples M1 M2 M3 M4 M5 M6 M7
Fig. 7. The AFM images of different membranes: (a) M1; (b) M2; (c) M5.
The relationship between contact angle and surface morphology is described by the Wenzel equation
(Wenzel, 1936). According to this model, if the surface is chemically hydrophilic, it will become even more
hydrophilic when surface roughness is added. If the surface is chemically hydrophobic, it will become even
more hydrophobic when surface roughness is added. From Fig. 7 (a), M1 has good roughness, which is in
good agreement with the following discussed SEM image. Fig. 7 (b) shows that coarse SiO2 particles increase
When dip-coating silica sols, the roughness of membranes increases first and decreases with the number
of dip-coating cycles. Since the PVDF membrane is coated by silica sols, it can be considered as a new
surface. For M3, the silica aerogel film is not formed and SiO2 nanoparticles adhere to the surface of PVDF,
causing pore blocking observed by SEM. Therefore, the roughness of M3 is lower than M1. The silica layer
contains many –OH groups and only a small amount of TMCS participates in the hydroxyl reaction with SiO2
nanoparticles. For M4~M6, the membrane surfaces are silica films with network structure as shown in SEM.
Therefore, the liquids can be well connected together on the surface protrusions so that the water droplets
adhere to the solid and air gaps explained by the Cassie-Baxter model (Cassie and Baxter, 1944). This means
that the increase in roughness reduces the contact area between the liquid and membrane surface. Therefore,
the roughness of M5 is the largest, with the largest contact angle as shown in SCA. The tapping-mode AFM
16
images of M5 is shown in Fig. 7 (c). For M7, the aerogel skeleton collapses due to fast solvent evaporation. A
large number of SiO2 nanoparticles stack together, so M7 has a larger roughness than M1. The silica layer
contains many –OH groups and it hardly participates in the hydroxyl reaction. All the results are consistent
Surface structures of the PVDF membranes are macro-porous with pore diameters from 200 nm ~ 1 μm
as indicated in Fig. 8 (a). However, in Fig. 8 (b) it can be seen that PVDF-SiO2 hybrid membrane has changed
from smooth to rough after 2.0% of hydrophobic SiO2 aerogel power is added. As verified by Fig. 8 (b), the
nanoparticles on the membrane surface are caused by agglomeration, which then results in pores reduction.
According to Fig. 8 (c)-(g), it is concluded that the PVDF/SiO2 composite bilayer membranes have different
pore structures. When the dip-coating cycles are limited, the silica sol on the surface is re-dissolved in EtOH.
So silica films are not formed, as shown in Fig. 8 (c). The more the solution is coated on the surface, the more
complete the silica films are formed. Especially in Fig. 8 (e), complete silica aerogel films with high porosity
are fabricated. The silica layer has network structure with an average pore diameter of about 10 nm. The
PVDF layer has macro-porous structure with a pore diameter of about 200 nm~1 μm. When performed with
more dip-coating cycles, the solvent evaporates rapidly. So silica aerogel film is cracked, as Fig. 8 (g) shows.
10 μm 2 μm 2 μm
为 为, 为
μ
,μ
500 nm 500 nm ,μ
500 nm
(g)
17
1 μm
Fig. 8. The SEM graphs of membranes: (a~g) M1~M7. Subscripts 1 and 2 represent the same kind of
SiO2 layer 6 μm
PVDF layer
20 μm
Fig. 9 shows cross-section of the membrane for M5. As seen, the meso-porous silica aerogel layer is
formed on the surface of PVDF layer. Continuous silica aerogel films are supported by PVDF layers to avoid
easy fragility. Silica aerogel is famous for its friability owing to porous nanostructure (Du et al., 2013). Herein,
the obtained PVDF/Si aerogel film has a moderate bend without cracking. The thickness of silica aerogel film
is 6 μm, which is consistent with that estimated by Eq. (1). It can be clearly seen that the PVDF layer is not
affected by the top layer and it has connected pores. The water vapor can permeate through these pores
efficiently.
Wide-angle X-ray diffractions in Fig. 10 are performed to study the membrane crystalline
18
M1
12000 M2
M3
10000 M4
Intensity (counts) M5
8000 M6
M7
6000
4000
2000
10 20 30 40 50 60 70 80 90
2-theta
Fig. 10. The XRD patterns for different membranes (M1, M2, M3, M4, M5, M6 and M7).
It can be seen that M1 has the two dominant peaks at 18.3°, 20.08° assigned to α(0 1 0), α(1 1 0),
respectively. It is well known that these diffraction peaks are characteristic of α-phase structure (Kim et al.,
2002). This means that PVDF membrane contains mainly α-crystal. When adding SiO2, XRD analysis shows
that the crystalline structure of M2 is not much changed, indicating there is no interaction between PVDF and
SiO2.
When silica sols are dip-coated on the surface of PVDF membrane, the diffraction peaks are not changed
obviously. For M3~M7, no obvious diffraction peaks of SiO2 aerogel films are obtained and the peaks of M5
at 24.30° are amorphous. The main reason is that SiO2 aerogel films have porous network structures and hence
no complete crystal structures exist. This can be seen more clearly in the next Fourier transform infrared
spectroscopy (FTIR).
19
3.6 M1
M2
3.2 M3
M4
2.8 M5
M6
Transmittance(%)
2915
840
2.4 M7
1232
2.0
1.6
1.2
0.8
Fig. 11. FTIR patterns of the membranes (M1, M2, M3, M4, M5, M6 and M7).
The infrared spectra is plotted in Fig. 11. We can see the changes of the chemical groups before and after
the dip-coating process of the silica films. Table 6 lists the wavenumbers of the FTIR peaks corresponded to
different groups and crystallites. There are main conclusions: (1) Confirmation of the crystalline structures
obtained by XRD; (2) Confirmation of the Si-O-Si and Si-CH3 groups (Yu et al., 2009). It can be obtained that
pure PVDF membrane contains mainly α-crystal. When adding SiO2, the PVDF-SiO2 hybrid membrane still
contains mainly α-crystal and it has no obvious asymmetrical Si-O-Si stretching. This may be explained that
drying in a oven leads to the collapse of SiO2 porous structure. When dip-coating with silica sols as listed in
Table 6, only M4~M6 has asymmetrical Si-O-Si stretching Si-CH3 stretching. It shows that only appropriate
Meanwhile, M1~M7 have β-phase characteristic peaks at 840 cm-1. It proves that the PVDF copolymer
possesses predominantly α-crystal crystalline structure and slightly β-crystal crystalline structure. Moreover, it
can be concluded that the crystalline structure of PVDF is not changed by dip-coating silica sols, indicating
that two layers have no chemical bonds and just stack together.
Table 6
3.5. Distribution of pore diameters of PVDF layer and SiO2 aerogel layer
To analyze the effects of pore size distribution on moisture transfer and heat conductivity, the structures
of M5 are measured. The two layers of M5 are made separately to measure the parameters. The distribution of
pore sizes of PVDF layer and SiO2 aerogel layer are characterized and plotted in Fig. 12.
1.0
Pore volume dV/dD (cm3/g/nm)
0.8
0.6
0.01
0.4
0.2
0.0
0.00
1 10 100 1000 10000 1 10 100
Pore diameter (nm) Pore diameter (nm)
Fig. 12. Distributions of pore diameters of (a) PVDF layer and (b) SiO2 aerogel layer.
Fig. 12 describes the distribution of pore sizes of the two layers. The PVDF layer is macro-porous with a
porosity of 77.76% and a mean pore size of 511.5 nm, allowing high vapor permeation. The SiO2 aerogel
layer is meso-porous with a porosity of 90.18% and a mean pore size of 6.6 nm. This small pore size can
effectively decrease heat conduction by limiting the molecular oscillations. The PVDF/Si aerogel composite
bilayer membrane can be manipulated to balance the paradox between permeability and heat conductivity.
21
3.6 Moisture transfer and heat conductivity
Heat and moisture transfer through membranes have been studied for several years in our laboratory and
they provide the theoretical basis for performance analysis (Zhang, 2006, 2008; Zhang and Li, 2017; Zhang et
al., 2011). Under constant mass flux boundary conditions, mass transfer coefficient ka is calculated by
For laminar flow, Nu and Sh are constant values. So the air side mass transfer coefficient ka is calculated
The total moisture transfer coefficient kt, the mean exchanged humidity between the two air flows Δωm
(kg/kg), the logarithmic mean humidity difference Δωlm, and the membrane mass transfer coefficient km are
calculated by
(4)
5
Moisture transfer resistance (m-1s) in the bilayer membrane composite membrane comprises two layers:
(5)
m m m
The PVDF membrane mass transfer coefficient km is 0.03289 m/s (the measurement from M1). When air
side resistance ka is constant, membrane side resistance km is the key factor for moisture transfer. For all the
composite membranes, the processes to prepare the porous PVDF membrane are the same. It can be assumed
that the mass transfer coefficient of PVDF layer km1 is fixed to 0.03289 m/s. Therefore, the mass transfer
The equivalent diffusion efficient of moisture in the membrane Dvm (m2/s) is calculated by:
δ (6)
m m
where δi is membrane thickness in ith layer (m) and Dvmi (m2/s) is the diffusion efficient of moisture in ith
layer, i=1-2. Mass transfer coefficient and moisture diffusivity in each part are listed in the Table 7.
Table 7
22
Mass transfer coefficient in each part and the total mass transfer coefficient for every membrane.
(μm) (m/s) (m2/s) (μm) (m/s) (m2/s) (m/s) (m/s) (m2/s) (g/(cm2s cmHg))
3.5
M1
M2
3.0 M3
M4
M5
2.5 M6
km (10-2m/s)
M7
2.0
1.5
1.0
0.5
0.05 0.10 0.15 0.20 0.25
V (m3/h)
Fig. 13. Effects of dip coating cycles on mass transfer coefficients through membranes themselves.
The mass transfer coefficients for membranes themselves are shown in Fig. 13. As seen, the mass transfer
coefficient of the PVDF is the largest among all the membranes. In other words, the moisture transfer
resistance of the PVDF is the smallest. For M3~M7, the membrane mass transfer coefficients increase at first
from M3 to M5. Then they decrease with increased dip-coating cycles from M5 to M7. As seen, the moisture
23
transfer coefficients of M5 and M2 appear very close. This means the mass transfer resistance of the
well-filmed PVDF/SiO2 composite bilayer membrane is close to that of the PVDF-SiO2 hybrid membrane.
From Fig. 13, the permeability of the PVDF membrane with a 150 μm thickness is 2.43×10-5 g/(cm2 s cm
Hg). The permeability of the PVDF-SiO2 membrane with a 150 μm thickness is 1.28×10-5 g/(cm2 s cm Hg).
The performance of the PVDF/SiO2 bilayer membrane with a 156 μm thickness is 1.15×10-5 g/(cm2 s cm Hg).
These values are a bit higher than an asymmetric cellulose triacetate hollow fiber membrane (0.72×10-5 g/(cm2
s cm Hg)) (Pan et al., 1978) and two orders of magnitudes higher than the TES liquid membrane
(1.44×10-7~2.16×10-7 g/(cm2 s cm Hg)) (Ito, 2000). Therefore, the PVDF/SiO2 bilayer membrane fabricated in
In the porous media, the gas diffusion theory considers three mechanisms: Poisseuille flow, ordinary
molecular diffusion, and Knudsen diffusion. (Khayet and Matsuura, 2004; Zhang, 2006, 2008; Zhang et al.,
2008). In our study for humidification with PVDF membranes, Kn≥10, therefore only Knudsen and ordinary
diffusion is considered.
In the composite membrane, there are two layers moisture transfer resistance (m2s kg-1):
tot m m (7)
where
δm (8)
m
ρa m
Ordinary diffusion coefficient Do of water vapor molecule in air is calculated as 2.52×10-5 m2/s by Ref.
(Tomaszewska et al., 2000), Knudsen diffusion coefficient Dk is calculated by Ref.(Tomaszewska et al., 2000)
T (9)
p
(10)
o o
The relational expression between equivalent diffusion efficient of moisture Dvm and effective diffusivity
24
m o
(11)
The diffusion coefficients calculated by Knudsen model and by the experimental are listed in Table 8.
Table 8
Parameter of M5 used in the analysis and diffusion coefficient calculated by Knudsen model.
PVDF layer 511.5 77.76 3.0 2.986 23.35 5.09×10-6 4.93×10-6 −3.26%
SiO2 aerogel layer 6.6 90.18 6.0 0.927 561.54 1.86×10-7 1.77×10-7 −5.18%
From Table 8, it can be clearly seen that the SiO2 aerogel layer has smaller pore sizes, higher porosity and
higher pore tortuosity, while the PVDF layer has larger pore sizes, lower porosity and lower pore tortuosity.
Compared the calculated and the experimental, the values of diffusion coefficient have a maximum difference
of 5.18%. The reason may be that pores are more zig-zagged. Therefore, the calculated value is relatively
There are two layer heat resistances in the composite membranes: resistance in the PVDF layer;
resistance in the silica layer. When the total membrane thermal conductivity is measured, the total heat
m m m (13)
For all the composite membranes, it can be assumed that the thermal conductivity of the PVDF layer is
equal to 0.08199 Wm-1K-1. Then the heat resistance of the PVDF layer is 1.8295 kW-1m2K. Therefore the heat
resistance of the silica aerogel layer can be calculated by Eq. (13). Thermal conductivity of each part and the
effective conductivity of the whole membrane are listed in Table 9. The ratios of heat resistance in each layer
25
to the total are plotted in Fig. 14.
Table 9
Thermal conductivity in each part and the heat resistance of the whole membrane.
3.5 rdm
rpm
3.0
2.5
rm (kW-1m2K)
2.0
1.5
1.0
0.5
0.0
M1 M2 M3 M4 M5 M6 M7
Membrane
Fig. 14. The ratios of heat resistance in each layer to the total heat resistance.
From Table 9, after doping with SiO2 aerogel powder, the thermal conductivity of the blend membrane
decreases from 0.08199 W m-1K-1 to 0.08078 Wm-1K-1 (a 1.5% decrease). After dip-coating with silica films,
the thermal conductivity of the membrane decreases from 0.08199 Wm-1K-1 to 0.04487 Wm-1K-1, significantly
(a 45.27% decrease). Therefore, the PVDF/SiO2 composite bilayer membranes are more promising than
PVDF-SiO2 hybrid membranes in thermal insulation. The heat conductivity of the PVDF/SiO2 bilayer
membranes (0.04487 Wm-1K-1) is also far lower than the PVDF/PDMS (0.071 Wm-1K-1) and
26
PVDF/PDMS-SiO2 (0.062 Wm-1K-1) membranes (Khayet and Matsuura, 2004) and it is close to the reported
When dip-coating silica sols for 60 cycles, the 150 μm PVDF support layer accounts for 52.62% of the
total resistance, and the silica thin film of 6 μm accounts for 47.38% of the total resistance, as shown in Fig.
14. The total heat resistance of PVDF/SiO2 composite bilayer membrane is up to 3.4767 kW-1m2K, which
increases by 87.23% upon M2 (only 1.8569 kW-1m2 K) and by 90.04% upon M1 (only 1.8295 kW-1m2K),
respectively. It can be obtained that the formed complete silica thin film has good thermal insulation.
Appropriate thickness of silica films can be designed and fabricated to balance the paradox between a lower
The mass transfer coefficient of PVDF layer is fixed at 0.03289 m/s. Substituting this value into Eqs.
(4)-(6), the relations between the SiO2 layer thickness δm2 and the membrane mass transfer coefficient km and
(14)
m
.4 5. 5δm
(15)
5 ( .4 5. 5δm )
The heat resistance of the PVDF layer is 1.8295 kW-1m2K. Substituting this value into Eqs. (12)-(13), the
relation between the SiO2 layer thickness and membrane heat resistance is obtained:
δm (16)
m . 5
. 4
The effects of SiO2 layer thickness on mass transfer coefficient and heat resistance are plotted in Fig.15.
27
16
2.5
3.5 km
rm 14
3.0
2.0
12
Pe' (10 g/(cm s cm Hg) 2.5
10
1.5
rm (kW m K)
2.0
km (10 m/s)
8
2
-1 2
-2
1.0 1.5
6
-5
1.0
(10.67,1.103) (10.67,4.76)
4
0.5
0.5 2
0.0 0.0 0
0 5 10 15 20 25 30 35 40
δm2 (μm)
Fig. 15. Effects of SiO2 layer thickness on mass transfer coefficient and heat resistance.
From Fig. 15, the greater the thickness is, the less the mass transfer coefficient is. The heat resistance
increases linearly with the thickness. It is worth noting that a larger thickness is not negatively related to a
lower permeability. Other parameters like pore diameter, porosity, tortuosity are also important parameters.
Therefore, in our study, other parameters are assumed to remain unchanged. When thickness is 10.67 μm, the
membrane mass transfer coefficient is 1.103×10-2 m/s, the permeability is 0.82×10-5 (g/(cm2 s cmHg)), and the
heat resistance is 4.76 kW-1m2K. This means a PVDF/SiO2 composite bilayer membrane with a silica layer
Therefore, a PVDF/SiO2 composite bilayer membrane with thermal insulation and acceptable
permeability is prepared. Meanwhile, the membrane is hydrophobic and mechanically durable. All these
characteristics make it possible to be widely used in the seawater desalination. In the future, the membrane
with thermal insulation and high permeability can be fabricated by improving other parameters like pore
4. Conclusions
A promising membrane is one that permits vapor permeation efficiently with low thermal conductivity.
For this respect, a dip-coating approach with the addition of DMF is taken to prepare PVDF/SiO2 composite
bilayer membrane with thermal insulation for seawater desalination. The optimum membrane is that
dip-coated silica sols with 60 cycles. With dip-coating silica sols on the surface of PVDF membranes, the
network structures are formed with high porosity. The results show that the heat conductivity decreases from
28
0.08199 Wm-1K-1 to 0.04487 Wm-1K-1, moisture permeability sacrifices from 4.93×10-6 m2/s to 2.42×10-6 m2/s.
The membranes made in this research are not only simple in method, easy accessible of raw materials, good in
low cost, low energy consumption, but also advantageous in membrane qualities like uniformity and
cracking-free. Although there is a slight drop in the mass transfer coefficient, the reduced heat losses is more
appreciable due to the tremendous decrease in thermal conductivity. This is the first time that heat
conductivity of membranes can be manipulated for seawater desalination. In the future, the PVDF/SiO2
composite bilayer membranes can also be alternatives for other membrane distillation processes, besides
Acknowledgement
This Project was supported by the National Science Fund for Distinguished Young Scholars of China, No.
51425601. It was also supported by the National Key Research and Development Program, No.
2016YFB0901404.
29
References
Adachi, T., Sakka, S., 1988. Sintering of silica gel derived from the alkoxysilane solution containing
Alkhudhiri, A., Darwish, N., Hilal, N., 2012. Membrane distillation: a comprehensive review. Desalination
287, 2-18.
Alves, V., Coelhoso, I., 2006. Orange juice concentration by osmotic evaporation and membrane distillation: a
Banat, F., Jwaied, N., Rommel, M., Koschikowski, J., Wieghaus, M., 2007. Performance evaluation of the
"large SMADES" autonomous desalination solar-driven membrane distillation plant in Aqaba, Jordan.
Banat, F.A., Simandl, J., 1998. Desalination by membrane distillation: a parametric study. Separation Science
Brinker, C., Hurd, A., Schunk, P., Frye, G., Ashley, C., 1992. Review of sol-gel thin film formation. Journal of
Brinker, C.J., Scherer, G.W., 1990. Sol-Gel Science: The Physics and Chemistry of Sol-Gel Processing.
Calabro, V., Jiao, B.L., Drioli, E., 1994. Theoretical and experimental study on membrane distillation in the
concentration of orange juice. Industrial & Engineering Chemistry Research 33, 1803-1808.
Cassie, A., Baxter, S., 1944. Wettability of porous surfaces. Transactions of the Faraday Society 40, 546-551.
30
Chen, T.-C., Ho, C.-D., 2010. Immediate assisted solar direct contact membrane distillation in saline water
Chen, Z., Rana, D., Matsuura, T., Yang, Y., Lan, C.Q., 2014. Study on the structure and vacuum membrane
Darmawan, A., Karlina, L., Astuti, Y., Sriatun, Motuzas, J., Wang, D.K., da Costa, J.C.D., 2016. Structural
evolution of nickel oxide silica sol-gel for the preparation of interlayer-free membranes. Journal of
Deshpande, R., Hua, D.-W., Smith, D.M., Brinker, C.J., 1992. Pore structure evolution in silica gel during
aging/drying. III. Effects of surface tension. Journal of Non-Crystalline Solids 144, 32-44.
Du, A., Zhou, B., Zhang, Z., Shen, J., 2013. A special material or a new state of matter: a review and
Efome, J.E., Baghbanzadeh, M., Rana, D., Matsuura, T., Lan, C.Q., 2015. Effects of superhydrophobic SiO 2
nanoparticles on the performance of PVDF flat sheet membranes for vacuum membrane distillation.
El-Bourawi, M., Ding, Z., Ma, R., Khayet, M., 2006. A framework for better understanding membrane
El-Dessouky, H., Ettouney, H., Al-Juwayhel, F., Al-Fulaij, H., 2004. Analysis of multistage flash desalination
Elimelech, M., Phillip, W.A., 2011. The Future of Seawater Desalination: Energy, Technology, and the
Geng, H., Wu, H., Li, P., He, Q., 2014. Study on a new air-gap membrane distillation module for desalination.
Godino, M., Pena, L., Rincón, C., Mengual, J., 1997. Water production from brines by membrane distillation.
Guillén-Burrieza, E., Blanco, J., Zaragoza, G., Alarcón, D.-C., Palenzuela, P., Ibarra, M., Gernjak, W., 2011.
Experimental analysis of an air gap membrane distillation solar desalination pilot system. Journal of
Hassan, A.M., Al-Sofi, M.A.K., Al-Amoudi, A.S., Jamaluddin, A.T.M., Farooque, A.M., Rowaili, A., Dalvi,
31
A.G.I., Kither, N.M., Mustafa, G.M., Al-Tisan, I.A.R., 1998. A new approach to membrane and thermal
seawater desalination processes using nanofiltration membranes (Part 1). Desalination 118, 35-51.
Hsu, S., Cheng, K., Chiou, J.-S., 2002. Seawater desalination by direct contact membrane distillation.
Ito, A., 2000. Dehumidification of air by a hygroscopic liquid membrane supported on surface of a
Jennings, S., 1988. The mean free path in air. Journal of Aerosol Science 19, 159-166.
Kaviany, M., 1991. Principles of Heat Transfer in Porous Media, second ed. Springer, New York.
Khayet, M., Matsuura, T., 2004. Pervaporation and vacuum membrane distillation processes: modeling and
Khayet, M., Mengual, J.I., Matsuura, T., 2005. Porous hydrophobic/hydrophilic composite membranes.
Kim, J.W., Cho, W.J., Ha, C.S., 2002. Morphology, crystalline structure, and properties of poly (vinylidene
fluoride)/silica hybrid composites. Journal of Polymer Science Part B: Polymer Physics 40, 19-30.
Lenza, R.F.S., Nunes, E.H.M., Vasconcelos, D.C.L., Vasconcelos, W.L., 2015. Preparation of sol-gel silica
samples modified with drying control chemical additives. Journal of Non-Crystalline Solids 423, 35-40.
Li, G.-P., Zhang, L.-Z., 2016. Investigation of a solar energy driven and hollow fiber membrane-based
Matz, R., Zimerman, Z., 1985. Low-temperature vapour compression and multi-effect distillation of seawater.
Mohamed, E.S., Papadakis, G., Mathioulakis, E., Belessiotis, V., 2006. An experimental comparative study of
the technical and economic performance of a small reverse osmosis desalination system equipped with an
Muzychka, Y.S., Yovanovich, M.M., 2009. Pressure Drop in Laminar Developing Flow in Noncircular Ducts: A
Ortiz, J., Expósito, E., Gallud, F., García-García, V., Montiel, V., Aldaz, A., 2006. Photovoltaic electrodialysis
system for brackish water desalination: Modeling of global process. Journal of Membrane Science 274,
138-149.
Pan, C., Jensen, C., Bielech, C., Habgood, H., 1978. Permeation of water vapor through cellulose triacetate
32
membranes in hollow fiber form. Journal of Applied Polymer Science 22, 2307-2323.
Pérez-Lombard, L., Ortiz, J., Pout, C., 2008. A review on buildings energy consumption information. Energy and
Qtaishat, M.R., Banat, F., 2013. Desalination by solar powered membrane distillation systems. Desalination
308, 186-197.
Tomaszewska, M., Gryta, M., Morawski, A., 1995. Study on the concentration of acids by membrane
Tomaszewska, M., Gryta, M., Morawski, A., 2000. Mass transfer of HCl and H 2O across the hydrophobic
Tu, K.L., Nghiem, L.D., Chivas, A.R., 2010. Boron removal by reverse osmosis membranes in seawater
Wang, P., Chung, T.-S., 2015. Recent advances in membrane distillation processes: Membrane development,
configuration design and application exploring. Journal of Membrane Science 474, 39-56.
Wenzel, R.N., 1936. Resistance of solid surfaces to wetting by water. Industrial & Engineering Chemistry 28,
988-994.
Wirth, D., Cabassud, C., 2002. Water desalination using membrane distillation: comparison between
system configured by a double-pass flat plate solar air heater. Desalination 205, 163-177.
Yu, L.-Y., Xu, Z.-L., Shen, H.-M., Yang, H., 2009. Preparation and characterization of PVDF–SiO2 composite
hollow fiber UF membrane by sol–gel method. Journal of Membrane Science 337, 257-265.
Yun, Y., Ma, R., Zhang, W., Fane, A., Li, J., 2006. Direct contact membrane distillation mechanism for high
Zhang, L.-Z., 2006. Fabrication of a lithium chloride solution based composite supported liquid membrane
and its moisture permeation analysis. Journal of Membrane Science 276, 91-100.
Zhang, L.-Z., 2008. A fractal model for gas permeation through porous membranes. International Journal of
Zhang, L.-Z., Li, G.-P., 2017. Energy and economic analysis of a hollow fiber membrane-based desalination
33
system driven by solar energy. Desalination 404, 200-214.
Zhang, L.-Z., Wang, Y.-Y., Wang, C.-L., Xiang, H., 2008. Synthesis and characterization of a PVA/LiCl blend
Zhang, L.Z., Niu, J.L., 2003. Laminar fluid flow and mass transfer in a standard field and laboratory emission
Zhang, X.-R., Zhang, L.-Z., Liu, H.-M., Pei, L.-X., 2011. One-step fabrication and analysis of an asymmetric
cellulose acetate membrane for heat and moisture recovery. Journal of Membrane Science 366, 158-165.
Zhao, D., Yang, P., Melosh, N., Feng, J., Chmelka, B.F., Stucky, G.D., 1998. Continuous mesoporous silica
films with highly ordered large pore structures. Advanced Materials 10, 1380-1385.
Zisman, W.A., 1964. Relation of the Equilibrium Contact Angle to Liquid and Solid Constitution, in: Fowkes
F.M., Contact Angle, Wettability, and Adhesion. VSP International Science Publishers, Los Angles, pp.
1-51.
Table Captions
Table 1. The effects of different viscosity and dip-coating speed on film quality.
Table 7. Mass transfer coefficient in each part and the total mass transfer coefficient for every membrane.
Table 8. Parameter of M5 used in the analysis and diffusion coefficient calculated by Knudsen model.
Table 9. Thermal conductivity in each part and the heat resistance of the whole membrane.
Figure Captions
Fig. 4. Effects of molar ratios of DMF to TEOS on the gelation time and viscosity of silicon sol.
34
Fig. 5. The films before and after doping DMF. (a) before; (b) after.
Fig. 7. The AFM images of different membranes: (a) M1; (b) M2; (c) M5.
Fig. 8. The SEM graphs of membranes: (a~g) M1~M7. Subscripts 1 and 2 represent the same kind of
Fig. 10. The XRD patterns for different membranes (M1, M2, M3, M4, M5, M6 and M7).
Fig. 11. FTIR patterns of the membranes (M1, M2, M3, M4, M5, M6 and M7).
Fig. 12. Distributions of pore diameters of (a) PVDF layer and (b) SiO2 aerogel layer.
Fig. 13. Effects of dip coating cycles on mass transfer coefficients through membranes themselves.
Fig. 14. The ratios of heat resistance in each layer to the total heat resistance.
Fig. 15. Effects of SiO2 layer thickness on mass transfer coefficient and heat resistance.
35
SiO2 aerogel layer
SiO2 PVDF
porosity: 90.18%
0.01
0.00
1 10 100
Pore diameter (nm)
0.6
0.4
0.2
0.0
16
2.5
3.5 km
14
3.0
rm
2.0
12
2.5
Pe' (10 g/(cm s cm Hg)
10
1.5
rm (kW m K)
2.0
km (10 m/s)
8
2
-1 2
-2
1.0 1.5
6
-5
(10.67,1.103) (10.67,4.76)
1.0 4
0.5
0.5 2
0.0 0.0 0
0 5 10 15 20 25 30 35 40
δm2 (μm)
Effects of SiO2 layer thickness on mass transfer coefficient and heat resistance.
A PVDF/SiO2 (Si aerogel) composite bilayer membrane of low heat conductivity is developed.
Measures are taken to solve the problem of cracking in forming SiO 2 layer.
A silica layer thickness of 10.67 μm has the largest heat resistance with an acceptable permeability.
The bilayer composite membrane is superior to SiO 2 blend membrane in thermal insulation.
36