Theoretical Modeling and Simulation PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 47

Theoretical modeling and simulation of AGMD and LGMD

desalination processes using a composite membrane

Item Type Article

Authors Im, Baek-Gyu; Lee, Jung Gil; Kim, Young-Deuk; Kim, Woo-Seung

Citation Im B-G, Lee J-G, Kim Y-D, Kim W-S (2018) Theoretical modeling
and simulation of AGMD and LGMD desalination processes using
a composite membrane. Journal of Membrane Science 565: 14–
24. Available: http://dx.doi.org/10.1016/j.memsci.2018.08.006.

Eprint version Post-print

DOI 10.1016/j.memsci.2018.08.006

Publisher Elsevier BV

Journal Journal of Membrane Science

Rights NOTICE: this is the author’s version of a work that was accepted
for publication in Journal of Membrane Science. Changes
resulting from the publishing process, such as peer review,
editing, corrections, structural formatting, and other quality
control mechanisms may not be reflected in this document.
Changes may have been made to this work since it was submitted
for publication. A definitive version was subsequently published
in Journal of Membrane Science, [, , (2018-08-07)] DOI: 10.1016/
j.memsci.2018.08.006. © 2018. This manuscript version is
made available under the CC-BY-NC-ND 4.0 license http://
creativecommons.org/licenses/by-nc-nd/4.0/

Download date 20/03/2023 02:10:57


Link to Item http://hdl.handle.net/10754/629414
Theoretical modeling and simulation of AGMD and LGMD desalination processes using a composite

membrane

Baek-Gyu Ima, Jung-Gil Leea,b, Young-Deuk Kimc, Woo-Seung Kimc,*

a
Department of Mechanical Engineering, Hanyang University, 222 Wangsimni-ro, Seongdong-gu, Seoul 04763,

Republic of Korea
b
King Abdullah University of Science and Technology (KAUST), Water Desalination and Reuse Center (WDRC),

Biological and Environmental Science and Engineering Division (BESE), Thuwal 23955-6900, Saudi Arabia
c
Department of Mechanical Engineering, Hanyang University, 55 Hanyangdaehak-ro, Sangnok-gu, Ansan,

Gyeonggi-do 15588, Republic of Korea

 Corresponding author. E-mail address: [email protected] (W.-S. Kim); Tel.: +82-31-400-5248; Fax: +82-

31-418-0153

Abstract

Most previous studies of air- and liquid-gap membrane distillation (AGMD and LGMD) processes using a

composite membrane have been focused on an experimental approach. In this paper, rigorous theoretical

investigations of the AGMD and LGMD processes were performed with a flat sheet type module using a

composite membrane comprised of a polytetrafluoroethylene (PTFE) active layer and a polypropylene (PP)

support layer. The model predictions were verified by comparing with measured data, where good agreement

between the prediction results and experimental data was obtained. It was observed that as the gap size increased

the AGMD permeate flux decreased exponentially with increased diffusion resistance. On the other hand, the

LGMD permeate flux decreased exponentially and then increased asymptotically after attaining a minimum at a

certain liquid-gap size (5 − 7 mm). This phenomenon was due to the onset and enhancement of a natural

convection, resulting in an improvement in heat and mass transfer in the liquid gap.

Keyword: Air-gap membrane distillation, Liquid-gap membrane distillation, Desalination, Heat and mass

transfer, Composite membrane


1. Introduction

Membrane distillation (MD) is a thermally-driven separation process in which only water vapors or other

volatile molecules move through porous hydrophobic membranes. Compared to conventional desalination

processes such as multi-stage flash (MSF), multi-effect distillation (MED), and reverse osmosis (RO), several

core advantages of the MD process are; (i) low sensitivity to salt concentration, (ii) theoretically 100% salt

rejection, and (iii) low operating temperature and pressure [1,2]. Such a MD process can be classified into: (i)

direct-contact membrane distillation (DCMD), (ii) air-gap membrane distillation (AGMD), (iii) vacuum

membrane distillation (VMD), and (iv) sweeping-gas membrane distillation (SGMD).

Among the MD configurations, DCMD has been the most widely studied due to its simple configuration, the

capability to produce a relatively high flux, and it has no additional equipment such as an external condenser to

condense water vapor, which is required in VMD and SGMD [3]. However, the DCMD process has the

drawback of a relatively higher heat loss over other MD configurations [4]. The heat loss is mainly caused by the

heat conduction through a single hydrophobic membrane as well as the heat losses to the ambient through the

module structure. Reduced heat loss in the MD process can be achieved by (i) increasing the pore size, porosity

and thickness of the membrane, and (ii) decreasing the thermal conductivity of the membrane. However, the

latter cannot be fulfilled as most hydrophobic polymer materials have very similar thermal conductivities.

Meanwhile, in the former case, there exists a conflict between low conductive heat loss (with a high membrane

thickness) and high mass transfer (with a low membrane thickness). In conclusion, it is difficult for a single

hydrophobic membrane to meet the requirements of an ideal membrane for the MD process, i.e., high mass

transfer and low conduction heat loss, with a single hydrophobic membrane [5]. Therefore, in order to avoid the

conflict mentioned above, a microporous hydrophobic/hydrophilic (or hydrophobic/hydrophobic) composite

membrane that consists of a thin active layer for high mass transfer and a thick support layer for low conductive

heat loss can be employed. Most studies on composite membranes have been experimentally carried out [5‒11],

while a few papers based on theoretical approaches have been reported [5,10,11]. Furthermore, the

aforementioned theoretical researches were mainly focused on the DCMD process using a composite membrane

and the developed models were based on a length-averaged lumped model, which disregards spatial variations of

both the feed and permeate flows [12].

The DCMD module configuration consists of two stream flows separated by a thin membrane with a

hydrophobic nature. However, this configuration is vulnerable to conductive heat losses. Among all of the

possible MD processes, one of the alternatives to cope with conductive heat loss is the AGMD, which constitutes
an air gap between the membrane and cooling plate, resulting in lower conductive heat loss and a higher thermal

efficiency. In addition, if a low temperature feed is used as a cooling stream, the latent heat can be recovered

through the condensation of the vapors on the cooling plate [3]. On the contrary, AGMD has a lower permeate

flux compared to DCMD due to the high mass transfer resistance in the air gap [13]. In this regard, liquid-gap

membrane distillation (LGMD) has been suggested to overcome the lower permeate flux and higher heat loss in

AGMD and DCMD, respectively. LGMD utilizes a liquid gap between the membrane and cooling plate, which

is typically filled with a liquid. Recently, Francis et al. [14] attempted to fill the gap with different types of

materials, such as a sponge (polyurethane), DI water, sand, and polypropylene mesh. It was found that the

permeate flux of LGMD is higher than that of AGMD at the same operating conditions. Furthermore, an

improved LGMD permeate flux was achieved due to an enhanced natural convection in the liquid gap with

increasing liquid gap thickness from 9 mm to 13 mm. Essalhi et al. [15] also carried out an experimental

investigation of AGMD and LGMD using a composite membrane. The permeate flux of LGMD was slightly

higher (2.2 − 6.5%) than that of AGMD at the feed inlet temperatures of 35 – 80 °C and an air/liquid gap of

4.532 mm. Therefore, there is a dearth of literature for both comprehensive experimental and theoretical studies

of the LGMD process.

In this study, theoretical investigations were performed to attain more comprehensive and systematic

understanding of the heat and mass transfer phenomena in the AGMD and LGMD processes using a commercial

hydrophobic microporous composite membrane with an active layer of polytetrafluoroethylene (PTFE) and a

scrim-backing support layer of polypropylene (PP). The main objective of this study was to develop the

integrated numerical analysis model of AGMD and LGMD processes, which consists of mass, momentum,

species, and energy balances of both the bulk feed and coolant flows coupled with the heat and mass transfer of

water vapor through the membrane and air/liquid gap. The theoretical model developed in this study was

validated with experimental data obtained from the literature [14,16]. Further studies were carried out to identify

the effects of the feed and coolant temperatures, the feed and coolant flow rates, surface porosity, and natural

convection on the permeate flux.

2. Theoretical models

The mathematical models for the AGMD and LGMD processes were formulated to include heat and mass

transfers. In both processes, a flat-sheet composite membrane is in direct contact with the feed stream, and an air

or liquid gap is interposed between the membrane and a condensation surface while a coolant flows in the
direction countercurrent to the feed. The mass and heat transfer mechanisms in the AGMD and LGMD processes

are illustrated in Fig. 1. For the simulation, the following assumptions are employed: (i) steady-state

incompressible flow, (ii) negligible heat loss to the ambient environment, (iii) only water vapor permeation

through membrane pores, (iv) atmospheric pressure in the air gap, and (v) unidirectional Newtonian fluid flow in

feed and coolant side. The SEM images of the composite membrane used in the present work are shown in Fig. 2

[12,14,16,41]. It is shown that the knot-fibril net structured PTFE active layer (dark gray) was partially covered

by the PP scrim-backing support layer (white gray).

Fig. 1. Schematic diagram of the heat and mass transfer mechanisms in (a) AGMD and (b) LGMD.
Fig. 2. SEM micrographs of a commercial PTFE/PP composite membrane. Clockwise from top left: 100×, 500×,

1,000× and 10,000× magnifications.

2.1. Air-gap membrane distillation (AGMD)

With the assumptions given above, the momentum, mass, species, and energy balances for the feed side can

be expressed in terms of pressure (Pf), velocity (vf), concentration (xf), and temperature (Tf) [1,4,17]. The

momentum balance is obtained by solving the z-component of Navier-Stokes equation as follows:

dPf 3 f
 2
vf (1)
dz hch

The overall molar balance and the molar species balance for the salt species yield Eqs. (2) and (3) in terms of

velocity (vf) and concentration (xf), respectively, as follows:

1 dv f vf  M M  dx f J
 2 s  w  (2)
X f dz X f  s  w  dz M f  sp hch

x f dv f v f M w dx f
 0 (3)
X f dz  w X 2f dz

The energy balance on the feed side yields the following equation on simplification:

d  f v f c p, f T f Qf
 (4)
dz  sp hch

For the coolant flow, the momentum and energy balances can be expressed in terms of the pressure (Pc) and

temperature (Tc), respectively, as follows [4]:


dPc 3
  2c vc (5)
dz hch

d c vc c p ,cTc Qc
 (6)
dz  sp hch

2.1.1. Heat transfer modeling

Heat transfer in AGMD includes the heat transferred across the feed boundary layer (Qf), composite

membrane (Qm), air gap (Qag), condensate film (Qcf), cooling plate (Qcp), and coolant boundary layer (Qc), as

shown in Fig. 3.

Fig. 3. Schematic of heat transfer through the composite membrane in AGMD.

The convective heat transfer across the feed boundary layer is written as follows [4,16,18]:


Q f  h f T f  T f ,m  (7)

where Tf is the bulk feed temperature and Tf,m is the interface temperature between the feed and membrane. As

turbulence promoter and membrane supporter, a non-woven net spacer is used to enhance the heat and mass

transfer in the feed channel. In order to incorporate an enhancement of heat and mass transfer by spacers, the

spacer factor (kdc) is employed by considering the structural parameters of spacer, as given in Eqs. (9) and (10).

Therefore, the convective heat transfer coefficient, hf, in the spacer-filled feed channel can be determined by the

modified Dittus-Boelter’s correlation by incorporating the spacer factor on the flat surface of a fully-developed

turbulent flow [2,16,19‒23].

h f dh
Nu   0.023kdc Re0.8 Pr 0.33 (8)
kf
0.039 0.086
 d spf  0.75    
kdc  1.654    sp sin    (9)
 hst    2 

 d spf
2
 sp  1  (10)
2lm sp sin 

Here, kf is the thermal conductivity of the feed solution, Nu is the Nusselt number, dh is the hydraulic diameter,

defined as 4 sp / 2  wch  hch  / wch hch  4 1    / d f  , Re is the Reynolds number, Pr is the Prandtl number, df

is the filament diameter of spacer, δsp is the spacer thickness, εsp is the spacer porosity, lm is the mesh size, θ is

the hydrodynamic angle, and wch is the channel width.

The overall heat transfer through the composite membrane can be calculated as follows [12]:

Qm  Qm,a  Qm,a  s (11)

where Qm,a and Qm,a‒s are the heat transfers across the active-only and active/support layers, respectively. Here,

the heat transfer through the active/support layer consists only of conductive heat transfer, as the mass transfer

across the active/support layer is negligible [12]. Therefore, the heat transfers can be described as follows:

 sf
Qm, a  J agmd H m, a 
Rm, a
T f ,m  Ta  (12)

1   sf
Qm,a-s 
Rm,a  s
T f ,m  Ts  (13)

where Ta and Ts are the interface temperatures between the active layer and the air gap and between the support

layer and the air gap, respectively, Jagmd is the mass flux, ∆Hm,a is the enthalpy of evaporation, and εsf is the

surface porosity, defined as the surface area of the PTFE active layer exposed to the permeate side divided by

the total membrane surface area. The thermal resistance, Rm, in Eqs. (12) and (13) can be described as follows

[12]:

a
Rm, a  (14)
km, a

   
Rm,a  s   a  s  (15)
 k m , a km , s 
 

 
where km,i   m,i kair  1   m,i ki , i = a, s. km,i is the effective thermal conductivity of the composite membrane

[12], kair is the thermal conductivity of air in the pores, and ki is the thermal conductivity of the membrane

material.
The heat transfers across the air gap are estimated as follows:

Qag ,a  J agmd H ag ,a   sf hag Ta  Tcf   (16)

  
Qag,a-s  1   sf hag Ts  Tcf  (17)

where Tcf is the interface temperature between the air gap and the condensate film and the heat transfer

coefficient across the air gap, hag, is expressed as follows:

kag Nu
hag  (18)
 ag

where Nu can be calculated by considering natural convection, Eq. (19), or conduction, Eq. (20), depending on

the geometry, configuration, material, and operating conditions as follows:

1/9 11  L /  ag  42
 L 
Nu L 
 0.197 Ra1/4 

, 0.5  Pr  2 (19)
  ag 
2 103  RaL  2 105

Nu  1 (pure conduction), RaL  2  103 (20)

with RaL 

g  Ts  Tcf  ag
3
 Pr (21)
 2

In the above equations, RaL is the Rayleigh number, g is the gravitational constant, β is the volume expansion

coefficient, ν is the kinematic viscosity, and L is the channel length.

Thus, the overall heat transfer through the air gap is given as shown below.

Qag  Qag ,a  Qag,a-s (22)

Also, the heat transfer across the condensate film is determined by the following equation [4,13,16,24]:


Qcf  hcf Tcf  Tcp  (23)

where Tcp is the interface temperature between the condensate film and the cooling plate, and the heat transfer

coefficient hcf can be calculated as follows [25‒27].

1/4
 g cf2 Hkcf3 
hcf  0.943   (24)
 Lcf Tcf  Tcp  
 

The heat transfer through the cooling plate is given by the equation shown below [4,13,16,24]:

Tcp  Tp 
kcp
Qcp  (25)
 cp
where Tp is the interface temperature between the cooling plate and the coolant, kcp is the thermal conductivity of

cooling plate material, and δcp is the thickness of the cooling plate.

The heat transfer across the coolant boundary layer is expressed as follows [4,13,16,24]:

Qc  hc (Tp  Tc ) (26)

where Tc is the coolant temperature and hc is the heat transfer coefficient at the spacer-filled coolant channel,

which is same as that defined in Eq. (8) at the spacer-filled feed channel.

The overall heat transfer in the AGMD process at steady state can be established as follows [4].

Q f  Qm  Qag  Qcf  Qcp  Qc (27)

Qm,a  Qag ,a , Qm,a s  Qag,a s (28)

2.1.2. Mass transfer modeling

Fig. 4. Schematic of mass transfer through the composite membrane in AGMD.

The permeate flux across the composite membrane is expressed as follows.


J a   sf Ca Pf ,m  Pa  (29)

As shown in Fig. 4, the relevant transport mechanisms for water vapor in AGMD are Knudsen diffusion and

molecular diffusion. The membrane distillation coefficient through the active-only layer can be estimated by the

equation below [16,18]:

1
 1 1 
Ca     (30)
 Ckn Cmol 

where Ckn is the Knudsen diffusion coefficient and Cmol is the molecular diffusion coefficient. The equations for
these coefficients are given as follows [4,16,18]:

4 d p a Mw
Ckn  (31)
3  a a 2 RT

Mw  a DP
Cmol  (32)
 P  Pv   a a RT
where εa is the active layer membrane porosity, D is the diffusion coefficient, P is the total pressure, dp is the

membrane pore diameter, τa is the active layer membrane tortuosity, δa is the active layer membrane thickness,

and R is the gas constant. The binary diffusion coefficient D is calculated by the Fuller-Shelltler-Giddings (FSG)

empirical equation expressed as follows [12]:

 T 1.75 
D  1.19 104  m  (33)
 P 
 m 

where Tm and Pm are the mean temperature and mean water vapor pressure, defined as Tm  (T f ,m  Ta ) / 2 and

Pm  ( Pf ,m  Pa ) / 2 , respectively.

In Eq. (32), the water vapor pressure without dissolved species in the water can be calculated using the

Antoine equation, as shown below.

 3816.44 
Pv  exp  23.1964 
T  46.13 
(34)

Due to the existence of dissolved species with molar concentrations at the feed side, the reduction of the vapor

pressure can be determined using Raoult’s law by assuming as an ideal solution for simplicity [4,12,28]:

Pf ,m  Pv (1  x f ,s ) (35)

where xf,s is the molar fraction of the solution and Pv is the water vapor pressure calculated by Eq. (34).

Also, the mass flux across the air gap is written as follows [4,18,29]:


J ag  Cag Pa  Pcf  (36)

where Pa is the interface pressure between the active layer and the air gap and Pcf is the interface pressure

between the air gap and the condensate film.

For the stagnant non-condensable gases trapped within the membrane pores and in the air gap between the

membrane and the condensing plate, the dominant transport mechanism is molecular diffusion expressed as

follows [4,18,29].
Mw DP
Cag  (37)
 P  Pv   ag RT
Both the permeate fluxes across the active-only layer (Ja) and the air gap (Jag) are assumed to be the same and

can be expressed as follows [4,16,18].

J agmd  J a  J ag (38)

Based on Eq. (38), Eqs. (29) and (36) can be rearranged, respectively, as follows.

J agmd Ja J agmd J ag
  Pf , m  Pa ,   Pa  Pcf (39)
 sf Ca  sf Ca Cag Cag

Combining the above equations yields the local permeate flux as follows.

1
 1 1 
J agmd  
  C C  P f ,m  Pcf  (40)
 sf a ag 

The mean permeate flux over the membrane length L is given by the following equation [1,12,30].

L

1
Jm  J agmd dz (41)
L 0

2.2. Liquid-gap membrane distillation (LGMD)

Theoretical models of LGMD are similar to those of AGMD described in the previous section. In particular,

the momentum, mass, and energy balance equations for the bulk feed and coolant flows in LGMD are the same

as those of AGMD, as defined in Eqs. (1)–(6). The main differences compared to AGMD models are the heat

and mass transfers through the composite membrane and the liquid gap, while the transport mechanisms for the

feed, condensate plate, and coolant are the same.

2.2.1. Heat transfer modeling

Fig. 5 illustrates the heat transfer mechanism in LGMD, which includes heat transferred across the feed

boundary layer (Qf), composite membrane (Qm), liquid gap (Qlg), condensate plate (Qcp), and coolant (Qc). In this

section, the heat transfers through the feed, condensate plate, and coolant are not described, as these are the same

as those of AGMD expressed in Eqs. (7), (25), and (26), respectively.


Fig. 5. Schematic of heat transfer through the composite membrane in LGMD.

The heat transfer through the composite membrane consisting of active-only and active/support layers can be

described as follows [1,12]:

 sf
Qm,a  J lgmd H a 
Rm,a
T f ,m  Ta  (42)

1   sf
Qm, a  s 
Rm, a  s
T f ,m  Ts  (43)

where Qm,a and Qm,a‒s are the heat transfers across the active-only and active/support layers, respectively. Here,

Rm,a and Rm,a‒s are calculated using Eqs. (14) and (15), respectively. The total heat transfer is then given as

follows [12].

Qm  Qm,a  Qm,a  s (44)

Additionally, in order to determine Ta in Eq. (42), the conductive heat transfer is considered as follows [12].

k 
Qs   w  Ta  Ts  (45)
 s 

The heat flux across the liquid gap is written as follows [31].

Qlg  hlg (Ts  Tcp ) (46)

The heat transfer coefficient, hlg, inside the gap filled with a liquid can be shown as follows:

klg Nu
hlg  (47)
 lg

where Nu can be obtained by considering natural convection or conduction, respectively, as follows [14,32,33]:
1  L /  lg  40
Nu  0.046 Ra1/3
L , 1  Pr  20 (48)
10  RaL  10
6 9

Nu  1 (pure conduction), RaL  2  103 (49)

The convective heat transfer coefficient in the range of 2×103 ≤ RaL < 106 is estimated by a linear

interpolation between Eqs. (48) and (49).

The overall heat transfer through the LGMD process at steady state can be established as follows:

Q  Q f  Qm  Qlg  Qcp  Qc (50)

Qm,a  Qs (51)

2.2.2. Mass transfer modeling

Fig. 6. Schematic of mass transfer through the composite membrane in LGMD.

The mass transfer of LGMD through the membrane has transport mechanisms such as Knudsen diffusion,

molecular diffusion, and viscous flow, as shown in Fig. 6. This phenomenon is the same as the mass transfer of

DCMD. The permeate flux through the composite membrane can be calculated as shown below [12].


J lgmd   sf Clg Pf ,m  Pa  (52)

The membrane distillation coefficient for the active layer is calculated using the following equation [12,34]:

1
 1 1 
Clg      C pf (53)
 Ckn Cmol 

where Ckn and Cmol are equal to the values calculated using Eqs. (31) and (32), respectively. Cpf is the viscous

flow coefficient [2,12,34], as shown below:


2
M w  a rp Pv
C pf  (54)
8  a a RT

where rp is the membrane pore radius and μ is the dynamic viscosity of water vapor.

2.3 Solution procedure

In both the AGMD and LGMD processes shown in Fig. 1, the bulk feed and coolant streams are circulated in

a countercurrent-flow manner. Thus, the velocities, concentrations, and temperatures are known at the inlets of

the feed and coolant solutions (i.e., z = 0 and L, respectively), while the pressures are known at their outlets (i.e.,

z = L and 0, respectively). For the simulations, the boundary conditions for the velocity, pressure, concentration,

and temperature at the bulk feed side are as follows:

V f ,in
v f (0)  v f ,in  , Pf ( L)  P0 , x f (0)  x f ,in , T f (0)  T f ,in (55)
 sp hch wch

where P0 depicts the ambient atmospheric pressure at the outlets of the feed and coolant channels, and the

subscript in depicts the inlets of the feed and coolant channels.

On the other hand, the boundary conditions for the velocity, pressure, and temperature at the coolant side are

as follows:

Vc ,in
vc ( L)  vc ,in  , Pc (0)  P0 , Tc ( L)  Tc,in (56)
 sp hch wch

In this study, the thermophysical properties of water and seawater were employed from [35]. The set of

coupled ordinary differential equations (ODEs) of the AGMD and LGMD processes, i.e., Eqs. (1)‒(6), including

both bulk feed and coolant flows and the characteristics of the membrane for heat and mass transfer, were

solved simultaneously with the boundary conditions given in Eqs. (55) and (56). The unknowns (i.e., Pf (0), vc

(0), and Tc (0) at z = 0) are guessed, which converts the boundary-value problem into an initial-value problem.

In this study, Broyden’s method, which is a quasi-Newton method for finding solutions of nonlinear equations,

was employed to solve the system of ODEs [1,12,17,37,38]. The solution for each control volume proceeds

forward until z = L, and then the presumed values are updated. Thus, this procedure is iterated until the

boundary conditions, Eqs. (55) and (56), are satisfied with an absolute error below 10-6.

3. Results and discussion

3.1. Validation of the numerical model


Fig. 7. Experimental (symbol) [14,16] and simulated (line) permeate fluxes in AGMD and LGMD with different

gap sizes (δgap = 9, 13 mm) at Tf = 40 − 80 °C, Tc = 20 °C, and Vf = Vc = 1.5 l/min

The numerical models developed for the AGMD and LGMD processes were evaluated for inlet feed

temperatures in the range of 40 °C to 80 °C while maintaining a constant coolant temperature and inlet feed and

coolant flow rates of 20 °C and 1.5 l/min, respectively. The feed salt concentration was kept constant at 4.2 wt%.

Fig. 7 shows the comparison of the experimental and predicted results of the permeate flux. All the experiments

were repeated to demonstrate the reproducibility of permeate fluxes measured at different feed water

temperatures, and the maximum deviations in the measured permeate fluxes were found to be less than 6% for

LGMD and 2% for AGMD, respectively [14,16]. It was shown that both the predicted and measured permeate

fluxes increased exponentially as the feed temperature increased. The model predictions were in good agreement

with the experimental results over the entire temperature range, with a maximum relative deviation of 10%.

3.2. Parametric analysis of AGMD and LGMD

Tables 1 and 2 show the specifications and characteristics of the spacer and composite membrane employed in

the present work, respectively. The surface porosity was confirmed to be 42% using CAD software based on the

Fig. 2. The membrane module specifications are listed in Table 3. The numerical model was assessed at various

operating conditions including the temperatures and flow rates of the feed and coolant streams and the

membrane surface porosity. Here, the feed salt concentration was assumed to be 4.2 wt%.
Table 1 Specifications of the spacer.

Material PP

Spacer thickness, δsp (mm) 0.8

Filament diameter, dspf (mm) 0.4

Angle between filaments, θ (deg) 90

Mesh size, lm (mm) 2

Table 2 Characteristics of the PTFE/PP composite membrane.


Material PTFE PP

Thickness, δ (m) 20 80

Porosity,  (%) 70 34

Mean pore size, rp (μm) 0.5 0.1

Liquid entry pressure, LEPw (kPa) 207 160

Table 3 Specifications of the membrane module.


Channel height, hch (m) 0.002

Channel width, wch (m) 0.05

Channel length, L (m) 0.1

Air/liquid gap thickness, δ (mm) 1 − 13

3.2.1. Effect of the feed temperature and flow rate

The results in Fig. 8 imply that the permeate flux increases with increasing feed temperature at a coolant

temperature of 25 °C for both AGMD and LGMD processes. A significant augmentation of the MD permeate

flux was achieved at the higher feed temperatures because of the enhanced transmembrane temperature

difference, i.e., MD driving force, as demonstrated previously [14,15,36]. In the AGMD processes considered in

this study, the Nusselt number (Nu) is always unity as the Rayleigh number (Ra) calculated by Eq. (21) is less

than 2×103. As a result, only conduction heat transfer will occur in the air gap, and based on Fourier’s law of

heat conduction, the larger the air gap thickness, the higher the heat transfer resistance. Therefore, the heat loss

through the air gap in the AGMD process can be ignored, mainly due to the very low thermal conductivity of air.
Meanwhile, the AGMD permeate flux decreases greatly with increasing air gap thickness, which is attributed to

increased diffusion resistance through the air gap [13,24,25,27]. It is also shown that the LGMD permeate fluxes

are significantly higher than those of AGMD at a given gap size, regardless of the feed temperature. This is due

to the nonexistence of additional diffusion resistance through the air gap and the higher transmembrane

temperature difference in LGMD. Similar to the trends of the AGMD permeate flux, as the liquid-gap size

increases from 1 mm to 5 or 7 mm, the LGMD permeate flux also decreases over the entire temperature range.

However, the further increase of liquid gap results in an asymptotic increase of the permeate flux. This

phenomenon occurs primarily due to the enhancement of natural convection, which results in a relatively higher

heat transfer across the boundary layer in the liquid gap leading to lower temperature polarization and thus, a

higher transmembrane temperature difference. In addition, at a fully-developed turbulent natural convection

regime (Ra ≥ 106) with a higher liquid-gap size, variations in the permeate flux tend to be negligible because the

influence of gap size on the permeate flux diminishes as Ra number increases [37,38]. These results agree well

with the experimental results [14] illustrated in Fig. 7, showing that the larger the liquid-gap size, the slightly

higher the permeate flux, but at higher feed temperatures above 70 °C the variance in the permeate flux is nearly

negligible.
Fig. 8. Effect of the feed temperature (Tf) in the range of 40 − 70 °C on (a) AGMD and (b) LGMD permeate

fluxes at Vf = 2.5 l/min, Vc = 1.5 l/min, Tc = 25 °C, and εs = 0.42.

Fig. 9 illustrates the effect of the feed flow rate ranging from 0.5 l/min to 5.0 l/min on the permeate flux of

both AGMD and LGMD with an inlet feed temperature of 70 °C, while keeping the other operating conditions

constant, as before. It is shown that the LGMD permeate fluxes are always greater than those of AGMD and

both MD permeate fluxes increase asymptotically with increasing feed flow rate, especially for all of the

LGMDs and AGMD with a 1 mm air-gap size. At lower feed flow rates, the boundary layer resistance on the

feed side is representative of the overall mass transfer resistance and thus, the permeate flux was largely

enhanced with increasing flow rate. This behavior can be attributed to the heat transfer augmentation through the

feed stream, resulting in a decrease of the temperature gradient across the feed-side boundary layer yet lower

temperature polarization [1,39‒41]. However, a further increase of the feed flow rate leads to a plateau in the

MD permeate flux profile as the overall mass transfer resistance is dominated by the diffusion (or heat transfer)

resistance in the air (or liquid) gap [42]. As the flow rate increases at the air and liquid gaps of 1 mm, the mass

transfer resistances in the gaps increase from 50% to 65% and from 46% to 77%, respectively, of the total mass

transfer resistance, whereas those across the feed-side boundary layers decrease from 30% to 9% and from 46%

to 12%, respectively. Meanwhile, at higher air-gap sizes (≥ 3 mm), the diffusion resistance in the air gap highly

limits the mass transfer for the AGMD process and thus, the effect of the feed flow rate on the permeate flux is

negligibly small.
Fig. 9. Effect of the feed flow rate (Vf) in the range of 0.5 − 5.0 l/min on (a) AGMD and (b) LGMD permeate

fluxes at Vp = 1.5 l/min, Tf = 70 °C, Tc = 25°C, and εs = 0.42.

3.2.2. Effect of the coolant temperature and flow rate

The influence of coolant temperature in the range of 10 ‒ 30 °C on both AGMD and LGMD permeate fluxes

while the other operating conditions mentioned previously were fixed is presented in Fig. 10. It is observed that

as the coolant temperature increases, the LGMD permeate flux decreases greatly as compared to that of AGMD.

This is because the interface temperature between the active layer and liquid gap, Ta, increases proportionally

with the coolant temperature due to the higher heat transfer through the liquid gap compared to the AGMD case,

resulting in a deterioration of the transmembrane driving force obtained from Eq. (52). In the AGMD process, on

the other hand, the permeate flux is not influenced by the coolant temperature because the heat and mass transfer
resistances in the air gap are predominant yet the interface pressure between the air gap and condensate film, Pcf,

calculated by Eq. (34) is not sensitive in the low temperature range from 10 °C to 30 °C.

Fig. 10. Effect of the coolant temperature (Tc) in the range of 10 − 30 °C on (a) AGMD and (b) LGMD permeate

fluxes at Vf = 2.5 l/min, Vc = 1.5 l/min, Tf = 70 °C, and εsf = 0.42.

Fig. 11 depicts the permeate fluxes of the AGMD and LGMD processes at different coolant flow rates. In

general, at a given coolant temperature, a sufficient cooling energy should be supplied to remove the latent heat

of condensation of the permeate vapor in both processes. It should be noted that a further supply of coolant to

both membrane modules can lead to a higher pumping power as a consequence of the higher pressure drop.

However, insufficient cooling and enhanced heat transfer resistance through the coolant-side boundary layer,
induced by a low coolant flow rate, may cause an increase of the interface temperatures between the air gap and

condensate film (Tcf) and between the active layer and liquid gap (Ta), respectively, for both the AGMD and

LGMD processes. Hence, the permeate flux will decrease. Above all, the most dominant factor in the permeate

fluxes of both AGMD and LGMD is the heat and mass transfer resistances through the air and liquid gaps.

Therefore, as shown in Fig. 11, the permeate flux variation of the AGMD process with the high gap resistances

is imperceptible to the coolant flow rate considered in this study whilst the LGMD permeate flux is dependent

on the coolant flow rate, especially at lower coolant flow rates.

Fig. 11. Effect of the coolant flow rate (Vc) in the range of 0.5 − 5.0 l/min on (a) AGMD and (b) LGMD

permeate fluxes at Vf = 2.5 l/min, Tf = 70 °C, Tc = 25 °C, and εs = 0.42.


3.2.3. Effect of the membrane surface porosity

Fig. 12 presents the permeate flux profiles with respect to the membrane surface porosity in the range of 0.1 –

0.7. As shown in Eqs. (40) and (52), the membrane surface porosity (εsf) is an important factor in determining

the MD permeate flux. It was found that with an increase of the membrane surface porosity, the permeate fluxes

increase asymptotically in both the AGMD and LGMD processes, which has already been demonstrated in the

DCMD process [12]. At larger air-gap sizes (≥ 3 mm), however, the higher mass transfer resistance through the

air gap is significantly more dominant than that of the membrane and thus, the permeate flux variations are

relatively insensitive even as the membrane surface porosity was increased [14,42]. In all cases of larger air-gap

sizes, the air-gap mass transfer resistances account for more than about 66% of the total mass transfer resistance,

while those across the membrane exhibit less than about 29%. On the other hand, the mass transfer at a 1 mm air

gap was significantly enhanced compared to that of larger air gaps, yet its magnitude becomes comparable to

that of the membrane and hence, the AGMD permeate flux is significantly influenced by the membrane surface

porosity. It is shown that as the surface porosity increases from 0.1 to 0.7 at the air gap of 1 mm, the mass

transfer resistances through the membrane decrease asymptotically from 51% to 13% of total mass transfer

resistance, whereas those in the air gap increase asymptotically from 39% to 69%.
Fig. 12. Effect of the surface porosity (εs) in the range of 0.1 − 0.7 on (a) AGMD and (b) LGMD permeate fluxes

at Vf = 2.5 l/min, Vc = 1.5 l/min, Tf = 70 °C, and Tc = 25 °C.

4. Conclusions

In the present work, rigorous theoretical models of AGMD and LGMD using a hydrophobic PTFE/PP

composite membrane were developed. The developed models involve mass, momentum, species and energy

balances in the feed and coolant streams, the mass and heat transfers through the membrane and air or liquid gap,

and the physical characteristics of the composite membrane. The effects of the feed and coolant temperatures,

the feed and coolant flow rates, and the membrane surface porosity on the MD performance were evaluated at

different air- and liquid-gap sizes.

It was noted that as the liquid gap increased the LGMD permeate flux decreased exponentially up to its

minimum at a certain liquid-gap size and then increased asymptotically. It was because the natural convection

that brings about an improvement of heat and mass transfer occurred at a certain liquid gap, and then was

enhanced with a further increase of the gap size. In general, the temperature polarization in the feed boundary

layer diminishes with an increase of the feed flow rate, leading to an increased interface temperature between the

feed and membrane and hence, the permeate flux. With larger air gaps, however, the AGMD permeate flux is

restricted by the high diffusion resistance in the air gap. The AGMD permeate flux is less affected by the coolant

temperature compared to the LGMD permeate flux. The negligible effect of the coolant temperature in AGMD is

due to the low sensitivity of the interface pressure between the air gap and condensate film at the lower

temperatures and also the heat and mass transfer resistances, which are strongly dominant in the air gap. The
effect of the coolant flow rate was marginal on the AGMD permeate flux because of the heat and mass transfer

resistances in the air gap, which are the most influential factors in the overall heat and mass transfers, although

the heat transfer resistance through the coolant-side boundary layer decreases as the coolant flow rate increases.

On the other hand, its effect on the LGMD permeate flux is relatively more significant at lower coolant flow

rates. As compared to the mass transfer resistance at larger air gaps, its magnitude at a 1 mm air gap is

comparable to that of the composite membrane. Therefore, the membrane surface porosity has a relatively large

influence on the AGMD permeate flux at a 1 mm air gap.

In all cases of the AGMD process, the permeate flux exhibits a maximum at an air gap of 1 mm. As expected,

the larger permeate flux is achieved with a smaller air gap with lower mass transfer resistance. Similarly, the

LGMD permeate flux at liquid gaps of less than 5 mm reveals a maximum at a gap of 1 mm, whereas above 5

mm the permeate flux increases asymptotically, leading to a maximum at a gap size of about 9 mm to 11 mm.

Acknowledgments

This work was supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP)

funded by the Ministry of Trade, Industry & Energy (MOTIE) of the Republic of Korea (No. 20153010130460

and No. 20174010201310).

Nomenclature

C Membrane distillation coefficient [kg/m2sPa]

Cp Specific heat capacity [kJ/kmolK]

D Diffusion coefficient [m2s]

dh Hydraulic diameter [m]

h Heat transfer coefficient [W/m2K]

hch Channel height [m]

g Gravity [m/s2]

∆H Enthalpy of water evaporation [J/kg]

J Permeate flux [kg/m2h]

k Thermal conductivity [W/mK]

L Channel length [m]

M Molecular weight [kg/kmol]


P Pressure [Pa]

Q Heat flux [W/m2]

Rm,a Thermal resistance for conduction through the active-only layer

Rm,a-s Thermal resistance for conduction through the active/support layers

T Temperature [°C]

V Volume flow rate [l/min]

v Velocity [m/s]

wch Channel width [m]

X Molar volume [m3/mol]

x Molar fraction of a solution [-]

Z Axial coordinate [m]

Dimensionless numbers

Nu Nusselt number [-]

Pr Prandtl number [-]

Re Reynolds number [-]

RaL Rayleigh number [-]

Greek letters

δ Thickness [mm]

ε Porosity [-]

θ Hydrodynamic angle [degree]

μ Dynamic viscosity [Pa s]

ρ Density [kg/m3]

τ Tortuosity [-]

ν Kinematic viscosity [m2/s]

Subscripts

a Active layer

ag Air gap
air Air

a‒s Active/support layer

c Coolant

f Feed

g Gas

lg Liquid gap

m Mean, membrane or mesh

p Membrane pore

s Support layer

sf Surface

sp Spacer

spf Spacer filament

v Vapor

w Water

Reference

[1] Y.-D. Kim, K. Thu, N. Ghaffour, K.C. Ng, Performance investigation of a solar-assisted direct contact

membrane distillation system, J. Membr. Sci., 427 (2013) 345–364.

[2] K.W. Lawson, D.R. Lloyd, Membrane distillation, J. Membr. Sci., 124 (1997) 1–25.

[3] L.M. Camacho, L. Dumée, J. Zhang, J.-D. Li, M. Duke, J. Gomez, S. Gray, Advances in membrane

distillation for water desalination and purification applications, Water, 5 (2013) 94–196.

[4] L.-H. Cheng, P.-C. Wu, J. Chen, Numerical simulation and optimal design of AGMD-based hollow fiber

modules for desalination, Ind. Eng. Chem. Res., 48 (2009) 4948–4959.

[5] M. Khayet, J.I. Mengual, T. Matsuura, Porous hydrophobic/hydrophilic composite membranes: Application

in desalination using direct contact membrane distillation, J. Membr. Sci., 252 (2005) 101–113.

[6] P. Peng, A.G. Fane, X. Li, Desalination by membrane distillation adopting a hydrophilic membrane,

Desalination, 173 (2005) 45–54.

[7] C. Feng, R. Wang, B. Shi, G. Li, Y. Wu, Factors affecting pore structure and performance of poly(vinylidene

fluoride-co-hexafluoro propylene) asymmetric porous membrane, J. Membr. Sci., 277 (2006) 55–64.

[8] R. Huo, Z. Gu, K. Zuo, G. Zhao, Preparation and properties of PVDF-fabric composite membrane for
membrane distillation, Desalination, 249 (2009) 910–913.

[9] J.A. Prince, V. Anbharasi, T.S. Shanmugasundaram, G. Singh, Preparation and characterization of novel

triple layer hydrophilic–hydrophobic composite membrane for desalination using air gap membrane distillation,

Sep. Purif. Technol., 118 (2013) 598–603.

[10] M. Qtaishat, M. Khayet, T. Matsuura, Guidelines for preparation of higher flux hydrophobic/hydrophilic

composite membranes for membrane distillation, J. Membr. Sci., 329 (2009) 193–200.

[11] D. Winter, J. Koschikowski, D. Düver, P. Hertel, U. Beuscher, Evaluation of MD process performance:

Effect of backing structures and membrane properties under different operating conditions, Desalination, 323

(2013) 120–133.

[12] J.-G. Lee, Y.-D. Kim, W.-S. Kim, L. Francis, G. Amy, N. Ghaffour, Performance modeling of direct contact

membrane distillation (DCMD) seawater desalination process using a commercial composite membrane, J.

Membr. Sci., 478 (2015) 85–95.

[13] G.L. Liu, C. Zhu, C.S. Cheung, C.W. Leung, Theoretical and experimental studies on air gap membrane

distillation, Heat Mass Transf., 34 (1998) 329–335.

[14] L. Francis, N. Ghaffour, A.A. Alsaadi, G.L. Amy, Material gap membrane distillation: A new design for

water vapor flux enhancement, J. Membr. Sci., 448 (2013) 240–247.

[15] M. Essalhi, M. Khayet, Application of a porous composite hydrophobic/hydrophilic membrane in

desalination by air gap and liquid gap membrane distillation: A comparative study, Sep. Purif. Technol., 133

(2014) 176–186.

[16] A.S. Alsaadi, N. Ghaffour, J.D. Li, S. Gray, L. Francis, H. Maab, G.L. Amy, Modeling of air-gap membrane

distillation process: A theoretical and experimental study, J. Membr. Sci., 445 (2013) 53–65.

[17] J.-G. Lee, W.-S. Kim, Numerical study on multi-stage vacuum membrane distillation with economic

evaluation, Desalination, 339 (2014) 54–67.

[18] A. Cipollina, M.G. Di Sparti, A. Tamburini, G. Micale, Development of a membrane distillation module for

solar energy seawater desalination, Chem. Eng. Res. Des., 90 (2012) 2101–2121.

[19] A.R. Da Costa, A.G. Fane, D.E. Wiley, Spacer characterization and pressure drop modelling in spacer-filled

channels for ultrafiltration, J. Membr. Sci., 87 (1994) 79–98.

[20] R.W. Schofield, A.G. Fane, C.J.D. Fell, Heat and mass transfer in membrane distillation, J. Membr. Sci., 33

(1987) 299–313.

[21] S. Al-Sharif, M. Albeirutty, A. Cipollina, G. Micale, Modelling flow and heat transfer in spacer-filled
membrane distillation channels using open source CFD code, Desalination, 311 (2013) 103–112.

[22] J. Phattaranawik, R. Jiraratananon, A.G. Fane, C. Halim, Mass flux enhancement using spacer filled

channels in direct contact membrane distillation, J. Membr. Sci., 187 (2001) 193–201.

[23] J. Phattaranawik, R. Jiraratananon, A.G. Fane, Effects of net-type spacers on heat and mass transfer in direct

contact membrane distillation and comparison with ultrafiltration studies, J. Membr. Sci., 217 (2003) 193–206.

[24] M. Khayet, T. Matsuura, Membrane Distillation, Elsevier, Amsterdam, 2011.

[25] A.M. Alklaibi, N. Lior, Transport analysis of air-gap membrane distillation, J. Membr. Sci., 255 (2005)

239–253.

[26] F.A. Banat, J. Simandl, Membrane distillation for dilute ethanol: Separation from aqueous streams, J.

Membr. Sci., 163 (1999) 333–348.

[27] S. Kimura, S.I. Nakao, S.I. Shimatani, Transport phenomena in membrane distillation, J. Membr. Sci., 33

(1987) 285–298.

[28] S.M. Shim, J.G. Lee, W.S. Kim, Performance simulation of a multi-VMD desalination process including the

recycle flow, Desalination, 338 (2014) 39–48.

[29] M. Khayet, Membranes and theoretical modeling of membrane distillation: A review, Adv. Colloid Interface

Sci., 164 (2011) 56–88.

[30] J.-G. Lee, W.-S. Kim, Numerical modeling of the vacuum membrane distillation process, Desalination, 331

(2013) 46–55.

[31] V.V. Ugrozov, I.B. Elkina, V.N. Nikulin, L.I. Kataeva, Theoretical and experimental research of liquid-gap

membrane distillation process in membrane module, Desalination, 157 (2003) 325–331.

[32] Y.A. Cengel, Heat transfer: A practical approach, 2nd ed., McGraw Hill, 2002.

[33] R.K. MacGregor, A.F. Emery, Free convection through vertical plane layers – Moderate and high Prandtl

number fluids, J. Heat Transfer, 91 (1969) 391‒401.

[34] L.-H. Cheng, P.-C. Wu, J. Chen, Modeling and optimization of hollow fiber DCMD module for desalination,

J. Membr. Sci., 318 (2008) 154–166.

[35] M.H. Sharqawy, J.H. Lienhard, S.M. Zubair, Thermophysical properties of seawater: a review of existing

correlations and data, Desalin. Water Treat., 16 (2010) 354–380.

[36] M. Essalhi, M. Khayet, Surface segregation of fluorinated modifying macromolecule for

hydrophobic/hydrophilic membrane preparation and application in air gap and direct contact membrane

distillation, J. Membr. Sci., 417-418 (2012) 163–173.


[37] N.C. Markatos, K.A. Pericleous, Laminar and turbulent natural convection in an enclosed cavity, Int. J.

Heat Mass Transfer, 27 (1984) 755‒772.

[38] K.J. King, Turbulent natural convection in rectangular air cavities, in: Queen Mary College, University of

London, UK, 1989.

[39] Y.-D. Kim, K. Thu, K.C. Ng, G.L. Amy, N. Ghaffour, A novel integrated thermal-/membrane-based solar

energy-driven hybrid desalination system: Concept description and simulation results, Water Res., 100 (2016) 7–

19.

[40] Y.-D. Kim, K. Thu, S.-H. Choi, Solar-assisted multi-stage vacuum membrane distillation system with heat

recovery unit, Desalination, 367 (2015) 161–171.

[41] J.-G. Lee, W.-S. Kim, J.-S. Choi, N. Ghaffour, Y.-D. Kim, A novel multi-stage direct contact membrane

distillation module: Design, experimental and theoretical approaches, Water Res., 107 (2016) 47–56.

[42] A.M. Alklaibi, N. Lior, Heat and mass transfer resistance analysis of membrane distillation, J. Membr. Sci.,

282 (2006) 362‒369.

Table captions

Table 1 Specifications of the spacer.

Table 2 Characteristics of the PTFE/PP composite membrane.

Table 3 Specifications of the membrane module.

Figure captions

Fig. 1. Schematic diagram of the heat and mass transfer mechanisms in (a) AGMD and (b) LGMD.

Fig. 2. SEM micrographs of a commercial PTFE/PP composite membrane. Clockwise from top left: 100×, 500×,

1,000× and 10,000× magnifications.

Fig. 3. Schematic of heat transfer through the composite membrane in AGMD.

Fig. 4. Schematic of mass transfer through the composite membrane in AGMD.

Fig. 5. Schematic of heat transfer through the composite membrane in LGMD.

Fig. 6. Schematic of mass transfer through the composite membrane in LGMD.

Fig. 7. Experimental (symbol) [14,16] and simulated (line) permeate fluxes in AGMD and LGMD with different

gap sizes (δgap = 9, 13 mm) at Tf = 40−80 °C, Tc = 20 °C, and Vf = Vc = 1.5 l/min.
Fig. 8. Effect of the feed temperature (Tf) in the range of 40−70 °C on (a) AGMD and (b) LGMD permeate

fluxes at Vf = 2.5 l/min, Vc = 1.5 l/min, Tc = 25 °C, and εs = 0.42.

Fig. 9. Effect of the feed flow rate (Vf) in the range of 0.5−5.0 l/min on (a) AGMD and (b) LGMD permeate

fluxes at Vp = 1.5 l/min, Tf = 70 °C, Tc = 25°C, and εs = 0.42.

Fig. 10. Effect of the coolant temperature (Tc) in the range of 10−30 °C on (a) AGMD and (b) LGMD permeate

fluxes at Vf = 2.5 l/min, Vc = 1.5 l/min, Tf = 70 °C, and εsf = 0.42.

Fig. 11. Effect of the coolant flow rate (Vc) in the range of 0.5−5.0 l/min on (a) AGMD and (b) LGMD permeate

fluxes at Vf = 2.5 l/min, Tf = 70 °C, Tc = 25 °C, and εs = 0.42.

Fig. 12. Effect of the surface porosity (εs) in the range of 0.1−0.7 on (a) AGMD and (b) LGMD permeate fluxes

at Vf = 2.5 l/min, Vc = 1.5 l/min, Tf = 70 °C, and Tc = 25 °C.


Table 1

Specifications of the spacer.

Material PP

Spacer thickness, δsp (mm) 0.8

Filament diameter, dspf (mm) 0.4

Angle between filaments, θ (deg) 90

Mesh size, lm (mm) 2


Table 2
Characteristics of the PTFE/PP composite membrane.
Material PTFE PP

Thickness, δ (m) 20 80

Porosity,  (%) 70 34

Mean pore size, rp (μm) 0.5 0.1

Liquid entry pressure, LEPw (kPa) 207 160


Table 3
Specifications of the membrane module.
Channel height, hch (m) 0.002

Channel width, wch (m) 0.05

Channel length, L (m) 0.1

Air/liquid gap thickness, δ (mm) 1 − 13


Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

You might also like