Sie sind auf Seite 1von 43

Accepted Manuscript

Title: Progress in Pressure Retarded Osmosis (PRO)


Membranes for Osmotic Power Generation

Author: Gang Han Sui Zhang Xue Li Tai-Shung Chung

PII: S0079-6700(15)00051-9
DOI: http://dx.doi.org/doi:10.1016/j.progpolymsci.2015.04.005
Reference: JPPS 927

To appear in: Progress in Polymer Science

Received date: 28-8-2014


Revised date: 5-3-2015
Accepted date: 13-4-2015

Please cite this article as: Han G, Zhang S, Li X, Chung T-S, Progress in Pressure
Retarded Osmosis (PRO) Membranes for Osmotic Power Generation, Progress in
Polymer Science (2015), http://dx.doi.org/10.1016/j.progpolymsci.2015.04.005

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.
Progress in Pressure Retarded Osmosis (PRO) Membranes for Osmotic Power Generation
Gang Han, Sui Zhang, Xue Li and Tai-Shung Chung*
Department of Chemical and Biomolecular Engineering, National University of Singapore, Singapore
117585

Abstract

t
The rapid increases in global energy consumption and greenhouse gas emissions have stimulated the

ip
exploration of renewable energy sources as alternative fuels. Osmotic pressure gradient energy released
from the mixing of water streams with different salinities is an unexploited resource of renewable

cr
energy. By employing a semipermeable membrane to control the mixing process, the osmotic pressure

us
gradient energy can be harvested in terms of electrical power via pressure retarded osmosis (PRO)
without causing adverse environmental impacts. The ideal of harvesting osmotic power via PRO was
proposed in the early seventies; however, the absence of effective membranes with desirable structure

an
and performance hindered further advancement of the PRO technology. During the last few years, a
significant progress in PRO technology has been achieved. Novel flat-sheet and hollow fiber polymeric
M
membranes with desired structure, mechanical robustness and permeation characteristics have been
developed for PRO applications. Membranes with a target power density of 5 W/m2 to produce
commercially viable PRO processes have been achieved. At this point of time, a comprehensive review is
d

imperative in order to summarize what we have accomplished and provide insights for the development
te

of next generation PRO membranes. After a brief introduction of the PRO process and the early PRO
development using the existing RO/NF and FO membranes, this review focuses primarily on novel and
p

the state-of-the-art PRO membranes. Furthermore, the requirements for fabricating effective PRO
ce

membranes will be discussed and future perspectives will be presented.


Ac

Keywords: osmotic pressure gradient energy, pressure retarded osmosis, polymeric membrane,
renewable energy, thin-film composite

*Corresponding author: chencts@nus.edu.sg; Tel: +65-65166645; Fax: +65-67791936

1
Page 1 of 42
Table of Contents
Nomenclature....................................................................................................................................
1. Introduction...................................................................................................................................
2. Theoretical Potential of Osmotic Pressure Gradient Energy ........................................................
3. Classification of Osmotic Processes .............................................................................................

t
ip
4. Basic PRO Principles for Osmotic Power Generation..................................................................
5. Power Density and Theoretical Power Generation Using PRO Membranes................................

cr
6. Chronicles of Membranes for PRO...............................................................................................

us
6.1 Early PRO work using RO membranes ..................................................................................
6.2 PRO performances of conventional FO membranes ..............................................................

an
7. Current Development of PRO Membranes...................................................................................
7.1 Integrally-skinned phase inversion and thin film composite (TFC) .......................................
M
7.2 TFC-PRO flat-sheet membranes.............................................................................................
7.3 Hollow fiber PRO membranes................................................................................................
7.4 Fouling and anti-fouling strategies for PRO membranes .......................................................
d

8. Conclusion and Perspectives.........................................................................................................


te

Acknowledgments.............................................................................................................................
p

References.........................................................................................................................................
ce

Figure Captions.................................................................................................................................
Tables................................................................................................................................................
Ac

2
Page 2 of 42
Nomenclature

PRO ═ pressure retarded osmosis


ICP ═ internal concentration polarization
IP ═ interfacial polymerization
ΔGmix ═ change in Gibbs energy (J/mol)
GB ═ Gibbs energies of the brackish water (J/mol)

t
ip
GS ═ Gibbs energies of the salt water (J/mol)
GF ═ Gibbs energies of the fresh water (J/mol)

cr
µ 0 ═ stand molar free energy (J/mol)

Vi ═ specific volume of component i (m3/mol)

us
R ═ gas constant (8.31441 J/mol K)
T ═ the absolute temperature (K)

an
xi ═ molar fraction of the component i (dimensionless)
z ═ the valence of an ion (equiv./mol)
M
F ═ the Faraday constant (96,485 C/equiv.)
∆ϕ ═ electrical potential difference (Volt)
c ═ salt molar concentration (mol/L)
d

V ═ volume (L)
te

i ═ the van’t Hoff factor (dimensionless)


ΔP ═ hydrostatic pressure difference (bar)
p

Am ═ membrane area (m2)


ce

Jw ═ water permeation flux (m s-1)


Js ═ reverse salt flux (g m-2 s-1)
W ═ membrane power density (W/m2)
Ac

A ═ water permeability coefficient (m s-1 bar-1)


B ═ salt permeability coefficient (m s-1)
πf ═ osmotic pressure of the bulk fresh water (bar)
πs ═ osmotic pressure of the bulk salty water (bar)
Δπbulk ═ the osmotic pressure difference between the bulk fresh water and salty water (bar)
Δπm ═ the osmotic pressure difference across the membrane (bar)
Δπeff ═ the effective osmotic pressure difference across the membrane selective layer (bar)
K ═ solute resistivity within the porous support layer (s m-1)

3
Page 3 of 42
S ═ membrane structural parameter (m)
t ═ thickness of the support layer (m)
τ ═ tortuosity of the support layer (dimensionless)
ε ═ porosity of the support layer (dimensionless)
Ds ═ solute diffusion coefficient in the support layer (m2 s-1)
σ ═ reflection coefficient (dimensionless)

t
k ═ mass transfer coefficient (m s-1)

ip
cr
us
an
M
d
p te
ce
Ac

4
Page 4 of 42
1. Introduction

Global energy consumption is projected to grow by 56% and the total energy usage will rise to 240 kilo
TWh in 2040 [1]. The demand for electrical power is fast-growing and its output will nearly double in the
next three decades [1]. Meanwhile, the reserves of conventional fossil fuels are decreasing and the
emission of greenhouse gases is changing the global climate. Facing the stringent environmental
regulations and high energy cost, unprecedented exploitation of alternative renewable energy sources

t
ip
has been observed [2].
Osmotic pressure gradient energy exploits the free energy released during the mixing of water
streams with different salinities. This energy could be harvested for power generation using membrane-

cr
based technologies such as pressure retarded osmosis (PRO) and reverse electrodialysis (RED) [3-5].
Compared to RED, PRO shows greater efficiency and higher power density and is more suitable to

us
extract power from high salinity gradients [5]. In a typical PRO process, water permeates spontaneously
across the semi-permeable membrane from the feed into the pressurized salt water because of the
chemical potential gradient across the membrane. The volume and hydrostatic pressure of the diluted

an
salt water are consequently increased which enables the generation of electricity when releasing the
pressurized water through a hydro-turbine [6-8]. It was estimated that the global osmotic power is in
the order of 1750–2000 TWh/year, and more energy is expected when concentrated brine and brackish
water are taken into account [9-11]. In addition, PRO is an environmentally-friendly technology because
M
no chemical and greenhouse gas are released during the power generation [11,12].
The conception and feasibility of PRO for osmotic power generation were recognized by
d

Norman and Loeb et al more than half a century ago [4,7,13,14]. However, the obtained energy
te

from the existing RO/NF membranes was far below the expectation due to unsuitable membrane

structures [15-17]. As a result, the development of PRO was suspended for a long time mainly
p

owing to a serious lack of effective membranes and incentive to find alternative energies.
ce

Statkraft of Norway was the first company that pioneered serious research on osmotic power and
Ac

built the first PRO prototype plant in 2009 by mixing river water and seawater across a semi-

permeable membrane [6,8,18,19]. This PRO prototype has provided a platform to facilitate

academic research and industrial exploration for new membrane development. Another hybrid

RO-PRO prototype was built by the ‘Mega-ton water system’ project in Fukuoka, Japan in 2010

[20]. This plant aimed to hybridize seawater reverse osmosis (SWRO) and waste water

reclamation systems for power generation. Based on Statkraft analyses, the minimum power

5
Page 5 of 42
density of flat-sheet and hollow fiber membranes for commercially viable PRO processes should

be at least 5 W/m2 and 3 W/m2, respectively [8,10,21]. However, the utilization of PRO for real

electricity generation is still economically challenging when comparing with fossil fuel because

of the lack of high performance PRO membranes which are the heart of PRO processes

t
ip
[10,11,22].

With the recent advances on the fundamental understanding of mass transports during osmotically

cr
driven processes and the design of various high performance PRO membranes, new perspectives have
been opened up for PRO processes and applications. Several reviews on PRO processes have been

us
reported [10,11,21-26]; however, none of them deal with membrane materials and fabrication methods
in details. Therefore, the purpose of this review is to focus on the material science and membrane
engineering of designing high performance PRO membranes. In addition to briefly introducing the

an
background of osmotic pressure gradient energy, PRO process and mass transports in PRO, and early
PRO studies using conventional RO/NF and FO membranes, the review focuses on the state-of-the-art
PRO membranes and their PRO performance as well as spearheads the future R & D directions.
M
2. Theoretical Potential of Osmotic Pressure Gradient Energy

Osmotic pressure gradient energy is the free energy released during the mixing of waters with
d

different salt concentrations [3-5]. The amount of energy available from the mixing of two solutions can
be theoretically calculated in terms of Gibbs energy from basic thermodynamics. For example, the free
te

energy available from mixing 1 m3 of salt water and 1 m3 of freshwater can be calculated as:
∆Gmix = GB − (GS + GF ) (1)
p

where ΔGmix (J/mol) is the change in Gibbs energy and GB, GS, and GF are the Gibbs energies of the
resultant brackish water, the salt water, and the fresh water (J/mol), respectively.
ce

Assuming the solutions are ideal, the chemical potential (µi) of component i in the solution can be
presented as [9]:
µi = µi0 + V i ∆P + RT ln xi + zi F∆ϕ
Ac

(2)

where µ 0 is the molar free energy under standard conditions (J/mol), Vi is the specific volume of
component i (m3/mol), ΔP is the pressure change compared to the atmospheric conditions (Pa), R is the
gas constant (8.31441 J/mol K), T is the absolute temperature (K), xi is the molar fraction of the
component i, z is the valence of an ion (equiv./mol), F is the Faraday constant (96,485 C/equiv.), and
∆ϕ is the electrical potential difference (Volt).
Since there is no pressure change and charge transport, the free energy difference can be
theoretically estimated from the chemical potential change of the system after and before mixing:
∆Gmix = ∑ (Gi ,S + Gi ,F − Gi ,B ) = RT ∑ [ci ,SVS ln(xi ,S ) + ci ,FVF ln(xi ,F ) − ci ,BVB ln(xi ,B )] (3)
i i

6
Page 6 of 42
where c is the molar concentration (mol/L) and V is the volume (L). The change in Gibbs energy during
mixing is negative since energy is released. Eq. (3) provides a good approximation for the theoretical
amount of free energy obtainable from the mixing of two solutions.
Fig. 1 shows the maximum energy that could be theoretically extracted from the mixing of freshwater
with different saline waters containing various NaCl concentrations [25]. The amount of free energy
could be released is strongly dependent on the salinity difference between the two mixing solutions. For

t
example, approximately 0.75 kWh of energy could be released when 1 m3 of freshwater flows into the

ip
ocean (assuming 0.55 M NaCl concentration), which means that 1 m3 s-1 of flowing freshwater can
potentially produce 2.7 MW power (1kWh=3.6 MJ). When the SWRO brine with a NaCl concentration of
1.1 M is purposely mixed with freshwater, 1.5 KWh m-3 of energy could be dissipated. The estimated

cr
global osmotic pressure gradient energy is in the order of 1750–2000 TWh/year, which is equivalent to
about one-half of the current annually hydropower generation [8,27]. In Europe, the equivalent energy

us
potential is estimated at 170 TWh/year [6,8]. In the United States, about 1700 km3 of freshwater flows
into the ocean every year which could potentially generate about 55 GW energy when assuming an
energy conversion efficiency of 40% [25]. In summary, the theoretical potential of osmotic pressure

an
gradient energy is very huge but how to effectively harvest it remains challenging.
Figure 1
3. Classification of Osmotic Processes
M
The osmosis phenomenon describes the spontaneous water transport across a semipermeable
barrier from a feed stream of higher chemical potential (i.e., low solute concentration) into another
stream of lower chemical potential (i.e., high solute concentration) [3-6,11-13,23]. The latter solution
d

with lower water chemical potential is termed as the draw solution, while the former one as the feed.
The semipermeable barrier generally represents a membrane that selectively allows water transport but
te

restricts the passage of solutes. During the osmosis process, the draw solution is rapidly diluted by the
permeate water, while the feed is concentrated. This process will continue until the chemical potential
p

across the membrane reaches the equilibrium. The osmotic pressure is defined as the hydrostatic
pressure that would stop the solvent (i.e., water) transport across the membrane when it is applied to
ce

the draw solution. The osmotic pressure (π) of a solution can be calculated via van’t Hoff equation as
[23,25]:
π = icRT (4)
Ac

-1
where i is the van’t Hoff factor, c is the molar concentration (mol L ), R is the universal gas constant
(8.31441 N m mol-1 K-1), T is the absolute temperature (K). Correspondingly, the osmotic pressure of
seawater at 25 °C is between 25 and 33 bar since the NaCl concentration varies from 3.0 to 4.0 wt. % in
different locations. The RO retentate from SWRO plants may contain salt ranging from 6.0 to 7.0 wt. %,
indicating higher osmotic pressures of 50-59 bar [25].
Fig. 2 describes three possible osmotic processes that happen between fresh water and salty water
partitioned by a semipermeable membrane [28]. The top nature occurring process is called forward
osmosis (FO) during which water spontaneously transports across the membrane driven by the osmotic
pressure gradient (Δπ) between the two solutions. When a hydrostatic pressure (ΔP) is applied to the
salty water side, the permeate water is retarded and even ceased when ΔP is equal to Δπ. At any stage
when ΔP is between 0 and Δπ, water still flows into the salty water because Δπ remains larger than ΔP.

7
Page 7 of 42
This phenomenon is termed as pressure retarded osmosis (PRO) where the net driving force for water
transport is reduced to Δπ−ΔP. When the trans-membrane pressure ΔP is greater than Δπ, the direction
of water permeation is reversed because water is forced to permeate through the membrane from the
salty water into the fresh water. This incident is referred to as reverse osmosis (RO) which has been
extensively used for seawater desalination. In principle, no extra energy is required for FO; energy could
be produced by PRO, while energy must be provided for RO.

t
Figure 2

ip
4. Basic PRO Principles for Osmotic Power Generation

The PRO process for osmotic power generation could be well explained via the above PRO osmosis

cr
phenomenon. Fig. 3 illustrates a schematic diagram of a typical PRO osmotic power process with a
continuous and steady state flow [19,29,30]. The feed streams have to be pretreated via various

us
filtrations in order to remove impurities and reduce membrane fouling. After that, the treated salty
water is pumped through a pressure exchanger before entering into the high pressure compartment at a
high hydraulic pressure (Ps). At the same time, pretreated fresh water is supplied to the other

an
compartment at no or low hydraulic pressure (Pf) on the other side of the PRO membrane. The salty
water faces the active layer of the membrane, while the fresh water flows against the porous membrane
support. With the aid of the driving force (Δπ−ΔP where ΔP=Ps−Pf) across the membrane, water
spontaneously transports through the membrane from the fresh water to the salty water at a flow rate
M
of ΔV.
∆V = J w × Am (5)
where Am is the membrane area and Jw is the water permeation flux.
d

Once the permeate water flows through the membrane, it results in an increase in pressure in the
te

high pressure compartment. In addition, the salty water is expanded with additional incoming volume
(ΔV) from the low pressure compartment. As a result, it is diluted to a new solution termed as brackish
water. The brackish water is split into two streams: one is to drive a hydro-turbine to produce electricity
p

and the other passes through the energy recovery device (i.e., pressure exchanger). The pressure
ce

exchanger transfers the pressure of the brackish water to the feed salty water, thus enabling a cost-
effective PRO system. The pressure exchanger must work efficiently in order to minimize any energy loss.
Otherwise, the energy generated from the turbine may not outweigh the energy required to pretreat,
pump and pressurize the feed waters [6,8,31]. Therefore, in addition to (i) advance membrane
Ac

performance, (ii) optimize module design and (iii) increase the salinity gradient between water streams,
system optimization is the 4th element necessary to improve the real power output of PRO plants [32].
Figure 3
5. Power Density and Theoretical Power Generation Using PRO Membranes

The membrane performance for PRO applications is usually evaluated in terms of power density W
which is defined as the power output per unit membrane area (W/m2). A PRO membrane with a high W
is essential because, to a great extent, it determines the required amount of membrane area and the
size of a PRO plant for a given capacity of energy production. Mathematically, W is determined by the
product of the trans-membrane hydrostatic pressure ΔP (bar) and the water permeation flux Jw (m s-1)
across the membrane [8,17].

8
Page 8 of 42
W = ∆P × J w (6)
Without considering the concentration polarization effects, the ideal Jw through a perfect
semipermeable membrane can be calculated as:
J w = A( ∆π − ∆P ) (7)
where A (m s-1 bar-1) is the intrinsic water permeability coefficient of the membrane, Δπ (bar) is the
osmotic pressure difference between the two solutions. Therefore, the ideal power output is:

t
W = A(∆π − ∆P)∆P (8)

ip
Mathematically, by differentiating Eq. (8) with respect to ΔP, the maximum power density (Wmax) occurs
when ΔP is equal to one half of the osmotic pressure difference (i.e., Δπ/2) across the membrane.

cr
( ∆π ) 2
Wmax = A × (9)
4

us
Therefore, the maximal theoretical power density in a PRO system is directly proportional to the
membrane water permeability A, and the square of the osmotic pressure difference. Operating PRO
processes at a pressure close to Δπ/2 is recommended in order to harvest the maximal power output.

an
For example, this ideal operating pressure is about 13.0-13.5 bar when using river water and seawater
solution pairing.
However, the effective osmotic pressure gradient across the membrane is less than the osmotic
M
pressure difference between the bulk salty water and fresh water (i.e., Δπ < πs-πf) in real PRO processes.
This is due to the detrimental effects of concentration polarization and reverse salt leakage.
Consequently, the water flux is dramatically decreased compared to the value calculated by Eq. (7).
There are two types of concentration polarization when the PRO membrane has an asymmetric
d

structure; namely external concentration polarization (ECP) and internal concentration polarization (ICP)
te

[16,17]. Fig. 4 illustrates the salt concentration and osmotic pressure profiles across an asymmetric
composite membrane operated in the PRO mode (i.e., the active layer facing the draw solution) [17]. Cs
and Cf are the salt concentrations of the bulk salty water and fresh water solutions, respectively. C1 is
p

the salt concentration at the interface between the salty water and the membrane selective layer. C1 is
smaller than Cs due to the dilution effects of the incoming water flux (Jw) from the fresh water, resulting
ce

in the dilution phenomenon called diluted ECP. Similarly, the salt concentration C3 at the interface of the
fresh water and the membrane is higher than Cf because of the depleted water and reverse diffused salt
(Js), leading to the phenomenon of concentrated ECP. The negative effects of the concentrative ECP on
Ac

fresh water are considered to be negligible when the support layer of the membrane is thick. The
combined effects of the diluted and concentrated ECP are the reduction of the osmotic driving force
from Δπbulk to Δπm. However, both ECP could be significantly mitigated by increasing the feed flow rates
on the external surfaces of the membrane.
Figure 4
On the other hand, the salts from the river water will migrate with the permeate water flowing across
the porous support and accumulate beneath the active layer. As a result, the salt concentration at the
interface between the membrane support and the active layer is much higher than that of the bulk fresh
water (C2>Cf). The enhanced salt concentration between the active layer and the support layer is termed
as concentrative ICP. Consequently, the effective osmotic pressure gradient across the active layer is
further reduced to Δπeff. It is worth noting that ICP usually causes more reductions in water flux and

9
Page 9 of 42
power density than ECP, especially at higher draw solution concentrations as discussed later. ICP could
not be reduced easily by increasing feed flow rate or inducing turbulence flow across the membrane
surface because it occurs inside the porous substrate. Studies have confirmed that ICP is generally
dependent on support layer morphology, solute characteristics and water permeation flux [6,33-35].
The solute resistivity K (s m-1) within the porous support layer is defined to characterize the ICP effects
as:

t
S
K= (10)

ip
Ds
where Ds is the solute diffusion coefficient in the membrane substrate (m2 s-1) and S is the structural

cr
parameter (m) that is a function of several membrane parameters:

S= (11)
ε

us
where t is the thickness of the support layer (m), τ is the tortuosity of the support layer (dimensionless),
ε is the porosity of the support layer (dimensionless).
In addition, since there are no perfect semipermeable membranes, salts from the draw solution will

an
back diffuse into the membrane substrate through the imperfect selective layer because of the
concentration gradient. This extra salt diffusion will exacerbate the negative effects of ICP and further
reduce the effective osmotic driving force across the membrane for PRO. Reflection coefficient (σ) has
M
been introduced to describe the ability of a membrane selective layer to preferentially allow water
permeation over salt permeation [36]. A perfect semipermeable membrane that rejects all the solutes
would have a reflection coefficient of 1. The reverse salt flux, Js (g m-2 h-1), permeating across the
d

membrane from the salty water to the fresh water can be described by [17]:
J s = B (C1 − C 2 ) (12)
te

-1
where B (m s ) is the salt permeability coefficient of the membrane active layer, C1 and C2 are the salt
concentrations at the interfaces of the active layer facing salty water and fresh water, respectively. The
p

reverse salt leakage also decreases C1 and π1; however, the net effects of Js on osmotic pressure gradient
across the membrane active layer may become detrimental when it is coupled with the ICP effects.
ce

Due to the effects of (1) ICP within the porous substrate, (2) ECP at the membrane surfaces and (3)
reverse salt flux, the effective osmotic driving force is much smaller than that between the bulk
solutions. As a result, the actual water flux Jact is written as follows:
Ac

J act = A(∆π eff − ∆P ) = A(π 1 − π 2 − ∆P ) (13)

When only considering the ICP and salt leakage that occur in the porous substrate, the salt
concentration at the interface between the substrate and the active layer can be expressed as [17,37]:
J act S B  J S 
C2 = C f exp( )+ (C1 − C2 ) exp( act ) − 1 (14)
Ds J act  Ds 
The first term on the left-hand side describes the ICP effect while the second term accounts for the
effect of salt concentration increase at the membrane interface due to the reverse salt flux into the
porous support.

10
Page 10 of 42
When only considering the ECP happening on the active layer facing the salty water due to the
permeate water from the other side (i.e., fresh water), the salt concentration at the active-saltwater
interface becomes:
J act B  J 
C1 = Cs exp(− )− (C1 − C2 ) 1 − exp(− act ) (15)
k J act  k 
where k is the mass transfer coefficient at the boundary layer. C1 consists of two terms. The first term

t
expresses the ideal salt concentration at the interface if B=0. In this case, C1 is equal to the product of

ip
the bulk salty water concentration Cs and an ECP factor of exp(-Jact/k). The second term accounts for the
salt loss due to the leakage across the membrane.

cr
As a result, the water flux across the membrane could be calculated as following using experimentally
measureable parameters and taking ICP, ECP and salt leakage into consideration [17,37-39]:
 

us
J J S
 π s exp(− act ) − π f exp( act ) 
 k Ds 
J act = A − ∆P  (16)
1 + B exp( J act S ) − exp(− J act ) 
 J act  

an
Ds k  
Thereby, the actual membrane power density (Wact) could be written as:
 J J S 
 π s exp(− act ) − π f exp( act )
M

 k Ds 
Wact = J act × ∆P = A − ∆P  × ∆P (17)
1 + B exp( J act S ) − exp(− J act ) 
 J act  Ds k 
 
d

where A is in m s-1 bar-1, Jact is in m s-1, πf and πs are in bar, B is in m s-1, ΔP is in bar, Wact is in W/m2, and k
te

is in m s-1 and Ds is in m2 s-1 and S is in m.


Clearly, the power density is a function of membrane water permeability (A), salt permeability (B),
and substrate structural parameter (S). Ideally, A is preferred to be as high as possible while B as small as
p

possible. However, an increase in A is always accompanied by an increase in B [40,41]. Therefore, one


ce

needs to find the best combination of A and B to maximize the power density. S must be small to
minimize ICP effects. To reduce ICP, previous experiences on forward osmosis (FO) membranes suggest
that a thin hydrophilic porous support layer with a high porosity and low tortuosity is preferred [42,43].
Ac

However, unless proper designs of PRO membrane materials and morphology, such membranes may
risk low mechanical properties and break at high pressure PRO tests [44,45].
Yip and coworkers have theoretically simulated the individual contributions of ICP, ECP and reverse
salt leakage to PRO performance [33]. As illustrated in Fig. 5, ICP, ECP and reverse salt leakage show
significant detrimental effects on water flux and membrane power density [33]. In the ideal case where
all the negative factors are absent (i.e., Δπeff =πs−πf), the water flux is around 50 L m-2 h-1 and the power
density is 18 W/m2 for a hypothetical PRO membrane using fresh water and seawater as feeds. However,
the actual performance is much lower than the ideal case. The calculated water flux and power density
dropped to 20 L m-2 h-1 and 6 W/m2, respectively, when taking these factors into calculations. A
significant difference between the theoretical and experimental power densities were also observed by
She et al [46] and it was attributed to the severe reverse salt diffusion across the PRO membrane and

11
Page 11 of 42
the aggravated ICP. The same work also found that the reverse salt flux increased with raising the
applied hydraulic pressure difference. Similar conclusions have been reported by Thorsen and Holt [8],
Zhang et al [39] and Sivertsen et al [47] who have experimentally and theoretically investigated the mass
transport, instant and accumulative effects of salt permeability on PRO processes.
Figure 5
In a word, an effective PRO membrane should possess an optimal combination of A, B, and S. The

t
membrane power density could be dramatically improved via directly tailoring the value of A, while it is

ip
constrained by the values of B and S [33,39,47]. Fig. 6 illustrates the iso-watt curves (i.e., constant power
density) as a function of A and B at different S values assuming the salt concentration of 28 g/L and a
membrane thickness of 50 μm at 20 ºC [47]. The iso-watt curves display similar trends for all S values. As

cr
aforementioned, an increase in S results in a reduction of power density. In addition, an increase in A
with a decrease in B within certain ranges will enhance the power density. However, the left-hand

us
vertical regions of the iso-watt diagrams indicates that a further decrease in B does not improve the
power density visibly, while the upper right-hand horizontal regions imply that a further increase in
water permeability does not augment power density significantly. Clearly, one must measure A, B and S

an
values of PRO membranes as a function of pressure and temperature, and then construct the iso-watt
diagram to identify the best operation window to run the PRO plant.
A comparison of Fig. 6 also indicates that membranes with a lower S value could achieve a higher
M
power density for a given pair of A and B [47]. Even though a smaller S could tolerate a higher B value by
enabling better diffusion of the accumulated salt in the porous substrate into the bulk feed solution and
minimize the ICP effect, the permeated salt will accumulate in the feed solution, increase its salinity, and
ultimately cause a reduction in power density of the overall system. Therefore, a small B value is
d

desirable to ensure a more stable salt permeability and a high water flux when increasing the
te

hydrostatic pressure [39]. In addition, the modeling results from Zhang et al [39] imply that the optimal
operation pressure to harvest energy in PRO processes can be greater than one-half of the osmotic
pressure gradient across the membrane if one can carefully design a PRO membrane with large A, small
p

B, and reasonably small S values. From an one-dimensional PRO mass exchanger model, Banchik et al
[48] also find that the optimal hydraulic pressure difference for practical PRO operations deviates
ce

significantly from one-half of the osmotic pressure difference as the dimensionless membrane area
increases and the ratio of flow rates between brine and feed water varies. A bigger membrane area is
required to produce a given power than the ideal case due to the effects of concentration polarization.
Ac

Figure 6
6. Chronicles of Membranes for PRO

6.1 Early PRO work using RO membranes

In the very early stages, commercially available RO/NF membranes originally designed for hydraulic-
pressure driven separation processes were tested in PRO for power generation [7,14-17,49-51]. Loeb
and Mehta were the pioneers using commercial Permasep RO membranes supplied by Du Pont for PRO
tests [7,14,15]. The membranes could withstand a high hydraulic pressure of 91.2 bar but showed
relative low peak power densities of less than 1.74 W/m2 at around 31 bar [7,14]. When both the
osmotic pressure gradient (Δπ) across the membrane and the operating hydraulic pressure became

12
Page 12 of 42
larger, a power density of 4.89 W/m2 at 50 bar was obtained by the same membranes [15,16]. FRL’s (i.e.,
the Fabric Research Lab) hollow fiber composite membranes consisting of a polysulfone support and a
furan outer skin were also tested by Loeb and Mehta [49]. This kind of membranes could achieve a
power density of 1.57 W/m2 at a hydraulic pressure gradient of 19 bar. Jellinek and Masuda [50] studied
the PRO performance of flat-sheet cellulose triacetate (CTA) RO membranes. The reported maximum
hydraulic pressure was about 17.2 bar and a power output of 1.62 W/m2 was achieved. Later, Mehta [51]

t
investigated the PRO performance of several spiral-wound and hollow fiber RO mini-modules.

ip
Reductions in water permeation coefficients were observed in PRO processes but no permanent
damage was found to these membranes after tests. The highest power density of 2.34 W/m2 was
obtained at about 20 bar by using a UOP CA/SW-3 spiral-wound mini-module.

cr
Fig. 7 summarizes the obtained power densities by using commercially available RO/NF membranes
in PRO processes. A power density of less than 1.22 W/m2 was obtained when seawater and fresh water

us
were used as the feeds where Δπ = 20-25 bar. A higher power density up to 4.89 W/m2 could be yielded
when more concentrated brine (Δπ >75 bar) was used. Nevertheless, these power densities or operation
conditions were far from applicability. The poor PRO performance was mainly due to the severe ICP

an
occurring inside the thick substrates of RO and NF membranes. Traditionally, the conventional RO
membrane has a very thick supporting layer of about 150-250 μm including the fabric and a hydrophobic
polysulfone layer in order to withstand very high pressures up to 100 bar. Not only did they retard the
M
free diffusion of ions, but also significantly reduce the effective osmotic pressure across the membranes
[17]. Consequently, this type of membranes is not suitable for PRO because it will significantly enhance
ICP and reduce PRO performance.
Figure 7
d

6.2 PRO performances of conventional FO membranes


te

Since both PRO and FO processes use similar osmotic principles and the desirable PRO and FO
membranes require similar characteristics such as high water flux, low salt reverse flux and minimal ICP
p

[11,23,26], several FO membranes were therefore tested in PRO [23,24,26,44-46,52-54].


The most investigated one is the asymmetric cellulose triacetate (CTA) based flat-sheet membrane
ce

produced by Hydration Technology Innovations (HTI, Albany, OR). As shown in Fig. 8, the CTA-FO
membrane has a thickness of around 50 μm. The hydrophilic nature of CTA ensures better wetting,
while the thin membrane thickness enables a much smaller structural parameter S than the
Ac

conventional RO membranes (i.e., 480 μm vs. 37500 μm) [11,22,52]. Achilli et al [24] reported the
maximum power density of 2.73 W/m2 and 5.06 W/m2 at 9.72 bar when using deionized water as the
feed and 35 g/L and 60 g/L NaCl solution as the salty water, respectively. She et al [46] found that the
CTA-FO membrane has a peak power density of 4.5 W/m2 utilizing 1 M NaCl as the draw solution and 10
mM NaCl as the feed. Kim and Elimelech [53] achieved a power density of 4.7 W/m2 using 2 M NaCl as
the draw solution and 0.5 M NaCl as the feed. Xu et al [54] evaluated a commercially available mini
Hydrowell® spiral wound FO (SWFO) module. The module was produced by HTI with a water
permeability of 2.2×10−12 m s-1 Pa-1 and has an active CTA membrane area of ~0.94 m2. When using 0.5
M NaCl as the draw solution and deionized water as the feed, a maximum power density of about 0.5
W/m2 was obtained. Fig. 9 summarizes the reported power densities of the CTA-FO membranes using
different feed schemes [6,19,23-26,46]. In most cases, the power densities of the HTI CTA-FO

13
Page 13 of 42
membranes are below the economically feasible value of 5 W/m2, although their PRO performances are
better than those of RO/NF membranes [6,8]. This is due to the fact that the HTI CTA-FO membrane is
hydrophilic and thin, but it has a relatively low intrinsic water permeability and a high salt permeability
(i.e., high reverse salt flux in PRO), as tabulated in Table 1.
Figure 8
Figure 9

t
Table 1

ip
In addition to the HTI CTA membranes, various other FO membranes developed in the last decade
were tested for PRO [11,23,26,55,56]. However, most conventional FO membranes were either

cr
deformed or damaged under high pressure PRO tests. Since FO processes require no or low-pressure
operation, FO membranes are normally designed to be very thin and porous in order to reduce

us
structural parameter and ICP. Therefore, they are not suitable for high pressure PRO tests because of
their poor mechanical properties. As shown in Fig. 10, Li et al [44] observed that the TFC-FO membranes
fabricated by the conventional approach were significantly compacted and deformed under high

an
pressures. The membrane with a sponge-like cross-section morphology is better than the finger-like one
to withstand high pressures and undergo less membrane deformation [44,45]. As a result, the power
densities obtained from experiments were well-below the theoretical prediction based on Eq. 17 using
M
the membrane transport characteristics determined by FO and RO tests [44]. Molecular designs of
membrane materials and morphology via optimization of polymer solutions and phase inversion
conditions with the aid of pre and/or post modifications must be advanced in order to improve
membrane strength and PRO performance.
d

Figure 10
te

7. Current Development of PRO Membranes

In the past five years, significant progresses have been made towards the PRO technology via
p

modeling and novel membrane design. The basic science and engineering to fabricate effective PRO
membranes has been developed and summarized as follows.
ce

7.1 Integrally-skinned phase inversion and thin film composite (TFC)


Ac

Two types of membrane preparation methods have been generally used for the preparation of PRO
membranes, including direct phase inversion and interfacial polymerization. The former is the most-
known route, involving phase inversion of a polymer dope in a non-solvent and the subsequent
formation of integrally-skinned membranes. On the other hand, interfacial polymerization route is used
to fabricate the thin film composite (TFC) membranes.
TFC membranes generally possess an asymmetric porous support and a top selective skin. The
microporous support provides the mechanical strength, while the selective layer performs the
separation. Interfacial polymerization has been widely used for the preparation of TFC membranes,
during which a cross-linked aromatic polyamide skin is formed on the substrate surface [57,58]. The
thickness of the polyamide layer is in the range of hundred nanometers due to the self-terminating
nature of the interfacial cross-linking reaction. The advantages of fabricating TFC membranes via

14
Page 14 of 42
interfacial polymerization are that the structure and properties of the substrate and the selective layer
can be individually tailored and optimized to achieve desired permeability and salt rejection.
Generally, the TFC-PRO membranes are fabricated by firstly preparing a porous membrane substrate
via phase inversion process, and then a polyamide selective layer is formed via interfacial polymerization
on top of the substrate. The morphology and mechanical properties of the microporous substrate are
particularly important because they directly determine the quality of the polyamide layer [59-62],

t
structural parameter, ICP effects and pressure tolerance under PRO tests [17,38,39,44,45]. So far, most

ip
TFC-PRO flat-sheet and hollow fiber membranes are produced by interfacial polymerization of m-
phenylene diamine (MPD) and trimesoyl chloride (TMC).

cr
7.2 TFC-PRO flat-sheet membranes

us
In this section TFC flat-sheet membranes recently developed for PRO osmotic power generation are
summarized and discussed. While maintaining a relatively small structural parameter to minimize ICP,
recent efforts have been mainly devoted to the enhancement of water permeability A during the

an
formation of polyamide layers and/or novel post-treatment processes. Various additives, such as bulky
monomers [63] and surfactants [64], have been included in reacting solutions to increase the intrinsic
free volume of the rejecting layer and water permeability. Post-treatments have also been employed to
increase the membrane water permeability including chlorine treatment [37,38], alcohol immersion
M
[44,45], and contact with DMF [64]. However, due to the simultaneous enhancements in salt
permeability, one has to carefully balance the two parameters to ensure a final net increment in power
density.
d

Li et al [63] designed several TFC-PRO flat-sheet membranes by manipulating the free volume of the
polyamide selective layer. A bulky monomer (i.e., p-xylylenediamine) was blended into MPD during
te

interfacial polymerization and a post-treatment of methanol immersion was conducted to swell up the
TFC layer. The former could enlarge and broaden the intrinsic free volume cavity of the polyamide layer,
p

while the latter could swell up polyamide chains, increase permeability, remove un-reacted monomers
and low molecular weight polymer chains [65-67]. They found that a moderate increment in free volume
ce

would significantly promote water permeability with a slightly decrease in salt rejection, and thus both
water flux and power density are increased. However, a too large increment in free volume would lower
both membrane selectivity and power density due to the effects of reverse salt flux and ICP as discussed
Ac

previously. Results from positron annihilation lifetime spectroscopy (PALS) on the polyamide layer
before and after PRO operations showed that the free volume of the TFC layer decreases and its
thickness becomes thinner due to high pressure compression. As a result, both water and salt
permeability increase because of a thinner TFC layer. However, the membrane PRO performance in
terms of power density and resistance against high pressures stays the same [63].
Han et al [38] developed another TFC-PRO flat-sheet membrane which consists of a polyamide
selective layer and a customized Matrimid® membrane support (see Fig. 11 (a) and (b)). The support
layer was designed to possess a fully sponge-like structure with robust mechanical strength. It can
withstand a hydrostatic pressure up to 15 bar. In order to enhance the transporting properties, the
polyamide layer was chemically modified using hypochlorite and then methanol. The modified TFC
membrane has higher water permeability but lower rejection. Under a mild post-treatment condition,

15
Page 15 of 42
the resultant membrane showed a power density range of 7 to 12 W/m2 at 15 bar when using various
synthetic water streams [38]. This good performance is resulted from a combination of (1) a robust
support layer with a small structural parameter and (2) a highly permeable (A=5.3 L m−2 h−1 bar−1)
polyamide active layer with a moderate salt permeability (B=2.0 L m−2 h−1). When further increasing the
degree of post-modification, the PRO performance decreased because the adverse effects of a higher
reverse salt flux coupling with ICP overwhelmed the benefit of an enhanced water permeability on PRO

t
performance [33,38,39].

ip
Figure 11
Since surfactants have been reported to improve the formation of the interfacially polymerized layer
of TFC membranes due to their amphiphilic nature [68-70], Cui et al [64] further improved the above

cr
TFC-PRO membranes via optimizing the interfacial polymerization reaction and alternative post-
treatments. By adding a certain amount of sodium dodecyl sulfate (SDS) in the MPD solution during

us
interfacial polymerization and post treating the resultant TFC membranes with N,N-dimethylformamide
(DMF), a power density of 18.09 W/m2 was achieved at 22 bar when using 1 M NaCl as the draw solution
and deionized water as the feed (see Fig. 12). This performance surpasses all flat-sheet PRO membranes

an
reported in literatures. Characterization data from PALS showed SDS could increase the free volume of
the polyamide layer without sacrificing the membrane selectivity. The DMF treatment was believed to
dissolve the loosely cross-linked parts of the polyamide selective layer without damaging the core part
M
[64].
Figure 12
Song et al [71] and Bui et al [72] recently reported a new type of TFC membranes supported by
customized nonwoven webs made by electrospun nanofibers. Electrospinning is a process driven by an
d

electrical potential to produce nano-polymeric fibers with diameters in the range of 40-2000 nm [73]. As
te

illustrated in Fig. 11 (c) and (d), the nonwoven nanofiber membrane has a super porous structure with
interconnected pores among nanofibers. Because of its high porosity and low tortuosity, it has a very
small structural parameter (i.e., S=150 μm) [71,72]. As a consequence, the detrimental effects of ICP
p

could be dramatically reduced. So far, the most efficient TFC-PRO membrane using nonwoven nanofiber
webs could achieve a power density of 15.2 W/m2 at 15.2 bar using synthetic brackish water (80 mM
ce

NaCl) and seawater brine (1.06 M NaCl) as the feed and draw solution, respectively [71]. However, the
membrane mechanical stability under high hydrostatic pressures in PRO is questionable at which the salt
leakage may be quite high. Hoover et al [35] incorporated an electrospun nanofiber backing support into
Ac

the substrate layer of the conventional TFC membrane to enhance membrane mechanical strength. The
membrane resistance to delamination under high fluid shear flow was improved which might make the
TFC membrane more suitable to sustain high PRO pressures. The successful utilization of electrospun
nanomaterials to improve membrane robustness and performance may boost the future research
interests in the exploration of new nanostructured PRO membranes.
HTI recently developed a TFC flat-sheet membrane for PRO applications which has a water flux more
than double the previous CTA membrane with a salt rejection of 99.3% [74,75]. The HTI TFC membrane
possesses a polysulfone porous support layer with an embedded woven mesh, and the total membrane
thickness is about 115 μm. Straub et al [75] studied its PRO performance via a lab-scale setup. Although
the woven mesh embedded in the support layer reinforces the membrane strength, the TFC membrane
may still be damaged under high pressures. As a result, a novel cross-flow membrane cell was designed

16
Page 16 of 42
in order to enhance its operation pressure for PRO applications. By using customized tricot fabric feed
spacers in the cell channels, the membrane could withstand a stable hydrostatic pressure of 48 bar for
more than 10 h and a power density of 14.1 W/m2 was achieved at 20.7 bar when using 1 M NaCl as the
draw solution and deionized water as the feed. The high performance is due to the optimized
membrane structure and permeation characteristics in addition to the novel membrane cell design.
In terms of membrane power density (W) and withstanding hydrostatic pressure difference (ΔP),

t
Table 2 summarizes the PRO performance of the recently developed TFC-PRO flat-sheet membranes

ip
under different test conditions. It is clearly found that effective TFC-PRO flat-sheet membranes have
been developed. Employing synthetic seawater water (0.59 M NaCl) as the draw solution and deionized
water as the feed, a high power density of 9.0 W/m2 could be achieved at around 13 bar [38]. When

cr
changing to more concentrated salty waters such as synthetic seawater brine (1.06 M NaCl), a power
density of 21.3 W/m2 could be obtained at 15.2 bar using deionized water as the feed [71]. The

us
advancements in PRO performance are mainly because of designing (1) the selective layer with a high
permeability but a relative low salt permeability and (2) the strong porous substrate with low resistance
for salt diffusion (small S) (see Table 3).

an
Table 2
Table 3
However, flat-sheet PRO membranes have several unavoidable shortcomings. In order to maintain
M
flow channel geometry and improve mass transfer near the membrane surface, a feed channel spacer is
needed for flat-sheet PRO membrane modules. The feed spacer will cause a hydraulic pressure loss in
the flow channel. The shadow effects from the spacer in the feed channel also reduce the permeation
flux across the membrane [76,77]. Moreover, the current feed spacers inevitably deform the PRO
d

membrane under high hydrostatic pressures (see Fig. 13 left) [77]. Membrane deformation not only
te

drastically reduces membrane selectivity but also increases its structural parameter. As a consequence,
the reverse salt flux and ICP effects are drastically increased, thus resulting in substantial reductions in
both water flux and power density as shown in the right-hand figures of Fig. 13 [77]. The extent of
p

membrane deformation is reported to be closely related to the membrane strength, applied hydraulic
pressure and spacer geometry [76,77]. Therefore, identification of spacers compatible with PRO
ce

membranes is of paramount importance for the development of effective flat-sheet PRO membrane
modules.
Figure 13
Ac

7.3 Hollow fiber PRO membranes

Hollow fiber membranes are tubular shaped membranes normally prepared via a non-solvent
induced phase inversion spinning process. In comparison to flat-sheet membranes, the hollow fiber
configuration has characteristics of higher surface area per module, self-mechanical support, and ease
of module fabrication [78]. Consequently, hollow fiber modules do not need feed spacers that make
them potentially suitable for PRO applications [47,79]. Not only could they minimize the membrane-
spacer interactions under high pressures, but also eliminate the aforementioned extra energy loss in the
feed flow channel of flat-sheet modules. Currently, both integrally-formed phase inversion dual-layer
membranes and TFC hollow fiber membranes have been explored for PRO power generation.

17
Page 17 of 42
7.3.1 Integrally-skinned PRO hollow fiber membranes

Integrally-skinned asymmetric hollow fiber membranes are prepared by the direct dry-jet wet phase
inversion of polymer dopes through the spinneret to the coagulant bath [80,81]. This type of
membranes features simplified and convenient fabrication processes. Such beneficial effects are
especially significant when outer-selective membranes are requested, as it takes much more efforts to
make outer-selective TFC membranes which will be discussed later. However, due to the constraints in

t
ip
material properties, the phase-inversion membranes usually have a lower water flux compared to TFC
ones.
The dual-layer configuration provides a possible solution to maximize membrane performance for

cr
PRO. Fu et al [80] are the pioneers to report the preparation of outer-selective dual-layer PRO hollow
fiber membranes by direct phase inversion. As shown in Fig. 14, the developed dual-layer hollow fiber

us
membranes consist of a PBI (polybenzimidazole)/POSS (polyhedral oligomeric silsesquioxane) outer
selective layer and a sponge-like PAN (polyacrylonitrile)/PVP (polyvinylpyrrolidone) inner support layer.
PBI is a promising candidate for PRO applications as it is hydrophilic and chemically stable. Owning to

an
the good mechanical properties and thermal stability, PAN is employed as the supporting substrate to
enhance the mechanical robustness of the hollow fibers. It was reported that the introduction of a small
amount of POSS nanoparticles (i.e. 0.5 wt. %) into the PBI outer layer had significant influences on both
the morphology and performance of the rejecting layer. In addition, the addition of PVP into the PAN
M
solution could minimize the delamination between the two layers due to its good compatibility with
PAN and PBI [80].
Figure 14
d

Later, the PBI/PAN dual-layer PRO hollow fiber membrane was further optimized by a novel post-
treatment method that involves flowing ammonium persulfate (APS) solution and water counter-
te

currently through the fiber to remove the PVP molecules entrapped in the substrate while keeping the
integrity of the interface [81]. As the APS concentration increases, the water flux in the PRO process was
p

increased while the salt leakage was slightly decreased. With the optimal APS concentration of 5 wt. %,
the post-treated membrane showed a maximum power density of 5.10 W/m2 at a hydraulic pressure of
ce

15.0 bar when using 1 M NaCl as the draw solution and 10 mM NaCl as the feed. To our best knowledge,
this is the best dual-layer PRO hollow fiber membrane directly fabricated from the non-solvent induced
phase inversion for osmotic power generation. However, there are still rooms to further improve the
Ac

PRO performance of these integrally-skinned hollow fiber membranes via (1) employing hydrophilic and
robust materials, and (2) well controlling the phase inversion process to further improve the membrane
mechanical strength, increase the membrane permeability and selectivity, but reduce the membrane
structural parameter.

7.3.2 TFC-PRO hollow fiber membranes

Recently, both inner- and outer-selective polyamide TFC-PRO hollow fiber membranes have been
successfully developed for osmotic power generation. Significant breakthroughs have been made in
terms of mechanical stability and water permeation flux.

18
Page 18 of 42
Han et al [82,83] designed a series of novel TFC hollow fiber membranes for PRO applications. The
newly developed TFC-PRO membranes consist of a polyamide selective skin formed on the lumen side of
the well-constructed Matrimid® hollow fiber substrates via interfacial polymerization [83]. Laboratory
PRO tests showed that the newly developed TFC hollow fiber membranes exhibited a power density as
high as 16.5 W/m2 with a very low specific reverse salt flux (Js/Jw) of 0.015 mol/L at a hydrostatic
pressure of 15 bar when using synthetic seawater brine (1.0 M NaCl) as the draw solution and deionized

t
water as feed [83]. In this work, new fabrication perspectives and design strategies were demonstrated

ip
to molecularly construct robust hollow fiber membrane supports. By manipulating the chemistry of
polymer solutions and the kinetics of phase inversion processes, hollow fiber supports with four
different micro-morphology and mechanical strengths were prepared, as illustrated in Fig. 15. It was

cr
observed that the strength of the microstructure of hollow fiber supports may determine the overall
membrane robustness rather than the apparent cross-section morphology. In addition, as shown in Fig.

us
16, pre-stabilization of the inner selective TFC hollow fiber membranes at high pressures could not only
improve the membrane water permeability with a slightly decrease in selectivity, but also reduce the
membrane structural parameter. The latter is probably attributed to the membrane expansion induced

an
by the high pressure applied in the lumen side. As elucidated in Fig. 17, a “critical pressure” was found
for each inner-selective TFC-PRO hollow fiber membrane, beyond which the polyamide selective layer
may experience irreversible changes, and significant defects will be formed [82,83]. The value of this
M
“critical pressure” is mainly dependent on the strength of the hollow fiber support. As a result, a robust
hollow fiber support should possess high stretch resistance and acceptable ductility. Effectively
manipulating phase separation kinetics during spinning using proper bore fluids and polymer solutions is
a promising strategy to maximize the mechanical strength of hollow fiber supports.
d

Figure 15
te

Figure 16
Figure 17
Chou et al [84] reported another TFC-PRO hollow fiber membrane with a polyethersulfone (PES)
p

substrate. The membrane showed a power density of 10.6 W/m2 when using 1 M NaCl as the draw
solution and 40 mM NaCl as the feed, but the membrane could only withstand a hydrostatic pressure of
ce

less than 10 bar [84]. Later, they further improved the fiber strength and PRO performance by using a
more robust material polyether-imide (PEI) as the substrate [85]. Meanwhile, the fiber diameter was
reduced and the cross-section morphology was tailored from a finger-like to fully sponge-like structure
Ac

(Fig. 18 top). The resultant TFC-PRO hollow fiber membrane could achieve a power density of 20.9 W/m2
at a pressure of 15 bar when using 1 M NaCl as the draw solution and 1 mM NaCl as the feed. However,
a relatively large specific reverse salt flux (Js/Jw) of 1.8 g/L was observed. In addition, the salt
permeability and structural parameter of the TFC hollow fiber membrane increased noticeably with an
increase in hydrostatic pressure imposed in the fiber lumen [85].
Figure 18
Li et al [86] prepared a series of P84 co-polyimide hollow fiber membrane supports with various
structures, dimensions, pore characteristics, and mechanical properties for inner-selective TFC-PRO
membranes by controlling the phase inversion process during spinning. The hollow fiber membrane that
was spun from a P84 co-polyimide/ethylene glycol (EG)/N-methyl-2-pyrrolidinone (NMP) solution with a
bore fluid of a water/EG/NMP mixture shows a small diameter and a high burst pressure up to 24 bar

19
Page 19 of 42
(see Fig. 18 bottom). The addition of NMP and EG into the bore fluid is believed to delay the demixing
and increases the interconnectivity of pores. They also found that the flow rate of draw solutions in the
lumen affected the PRO membrane. A high flow rate may increase the local pressure and shear the inner
wall of hollow fibers, leading to structural deformation of the TFC membranes [86].
Recently, Zhang et al [39,87] successfully developed a high performance TFC-PRO hollow fiber
membrane with ultra-high power density and robustness by using a molecularly engineered

t
polyethersulfone (PES) hollow fiber membrane substrate. As displayed in Fig. 19, the newly developed

ip
PES hollow fiber comprises 3 layers: (1) a highly asymmetric and porous outer layer, (2) a layer of short
macrovoids near the inner skin, and (3) a relatively dense thin sponge-like layer with small surface pores,
suitable for good deposition of the polyamide layer [39]. The highly porous outer layer and macrovoids

cr
may reduce ICP, while the relatively dense skin layer underneath the TFC layer help maintain good
mechanical stability and stress dissipation. As a result, the newly developed TFC-PRO hollow fiber

us
membrane could produce a maximum power density of 24.0 W/m2 at 20.0 bar with a very low specific
reverse salt flux of less than 1 g/L using 1 M NaCl as the draw solution and deionized water as the feed
(see the sample M in Fig. 20). So far, this is the best PRO membrane in terms of withstanding pressure

an
and power density. The importance of a low reverse salt flux has also been emphasized in this work.
Consistent with modeling results, a large salt permeability (B) not only causes an instant drop in the
initial water flux but also promotes the flux decline at high hydraulic pressures. Thus, both the optimal
M
operating pressure and maximal power density are reduced. Furthermore, it was found that a high B
could cause significant salts accumulation in the feed along the large membrane module, leading to
large reductions in both water flux and power density [39].
Figure 19
d

Figure 20
te

Table 4
Table 4 summarizes the permeation properties (A, B, S) and maximum power density (Wmax) of the
up-to-date TFC-PRO hollow fiber membranes. Effective TFC-PRO hollow fiber membranes with proper
p

membrane structure, balanced permeation properties, and high mechanical strength have been
demonstrated. The best membrane can withstand a hydrostatic pressure (ΔP) larger than 20 bar and
ce

harvest a power density of 24 W/m2 [87]. The newly developed TFC-PRO hollow fiber membranes show
great potential for osmotic power generation.
In summary, breakthroughs on high performance inner selective TFC-PRO hollow fiber membranes
Ac

depend not only on the improvement of polyamide layers, but also largely on the enhancement of the
mechanical strength of hollow fiber substrates. Theoretically, the optimal pressure on the salty water
side for the maximum power density is equal to one-half of the osmotic pressure difference across the
membrane (see Eq. (9)). However, the operating pressure in PRO is normally constrained by the limited
mechanical strength of the PRO hollow fiber membranes. Fig. 21 shows the polymer materials which
have been employed for the fabrication of effective hollow fiber supports thus far. They are all
intrinsically robust materials consisting of mechanically strong benzene rings. In addition to the careful
selection of the membrane material, the microstructure across the hollow fiber substrate should be
carefully designed with a balanced asymmetry. Pointed out by Han et al [83] and Zhang et al [87], the
desirable hollow fiber substrate for high performance TFC-PRO membranes should have a highly porous
support layer to reduce internal concentration polarization, while a relatively dense cushion layer

20
Page 20 of 42
beneath the polyamide layer is needed to re-distribute the stresses and stabilize the membrane, as
illustrated in Fig. 22 [87]. Table 5 lists the characteristics of the porous hollow fiber substrates for the
reported inner-selective TFC-PRO membranes. The pore size on the inner surface should be small with a
narrow pore size distribution to ensure the mechanical stability of the selective layer under high
pressures. In addition, the microstructure across the hollow fiber should be intimately interconnected to
have high fracture resistance that can effectively dissipate the stresses from the high pressure water

t
[83,87]. The fiber dimension and wall thickness also significantly influence the strength and performance

ip
of the TFC-PRO membranes. A critical wall thickness was observed by Zhang et al [39] to ensure
sufficient mechanical stability of the TFC-PRO hollow fiber membranes and thus a low salt permeability
at high pressures.

cr
Figure 21
Figure 22

us
Table 5

In comparison with inner-selective membranes, the outer-selective hollow fiber shows a lower

an
pressure drop in the draw solution side for the actual PRO applications. In addition, the outer-selective
hollow fiber provides more surface area per module. However, most of the reports on PRO hollow fiber
membranes have been focused on inner-selective membranes due to the difficulties of conducting
M
uniform interfacial polymerization on the outer surface. Recently, Sun et al [88] overcame the
challenges and developed a novel process to fabricate outer-selective TFC-PRO hollow fiber membranes.
By applying a vacuum-assisted interfacial polymerization, a defect-free polyamide selective layer was
successfully formed on the outer surface. As shown in Fig. 23, the polyamide layer shows the typical
d

“ridge-valley” morphology. The PRO performance was further improved by optimizing the pore size and
mechanical strength of the Matrimid® hollow fiber support and then modifying the polyamide selective
te

layer with a polydopamine coating. The newly developed membranes can withstand 20 bar with a peak
power density of 7.63 W/m2 when using 1M NaCl as the draw solution and deionized water as feed. This
p

PRO performance is equivalent to 13.72 W/m2 of its inner-selective hollow fiber counterpart with the
same module size, packing density, and fiber dimensions [88]. This study may provide insightful
ce

guidelines for the preparation of outer-selective TFC membranes for PRO applications.
Figure 23
Ac

7.4 Fouling and anti-fouling strategies for PRO membranes

Similar to other membrane-based separation processes, membrane fouling is always a challenge for
PRO membranes which leads to a severe decline in water flux and power density. Due to the
detrimental effects of concentration polarization, the reverse diffusion of solutes and the high
hydrostatic pressures applied on the draw solution, the fouling behaviors of PRO membranes are very
significant and complicated [89-94]. Zhang et al [90] found that gypsum scaling in PRO was strongly
affected by the chemistry of draw solutions and their concentrations, the applied hydrostatic pressure
and the membrane material and structure; in addition to the saturation index of the bulk feed solution.
She et al [91] and Chen et al [92] noticed that organic fouling is exacerbated by the presence of divalent
cations (i.e., Ca2+ and/or Mg2+) in the draw solution due to the reverse salt diffusion. Also, increasing

21
Page 21 of 42
hydrostatic pressure during PRO operations shows complex impacts on fouling behaviors as a result of
competing effects of increased reverse salt diffusion and reduced water flux. Chen et al [93] reported
that the co-existence of gypsum crystals and alginate would lead to more severe combined fouling than
the sum of individual ones and thus resulted in a greater water flux decline under 0 bar. However, such
gypsum-alginate synergistic fouling was not observed under high hydrostatic pressures in PRO which
might because that the increased reverse salt diffusion will inhibit the formation of gypsum crystals.

t
Thelin and coworkers [94] observed that the accumulated loading of natural organic matter (NOM)

ip
greatly affected the flux decline in PRO. The same study also found that cellulose acetate (CA)
membranes exhibited much less fouling than TFC membranes due to their higher surface hydrophilicity.
Thorsen et al [8] observed that the PRO fouling propensity increased with increasing the water

cr
permeation flux and feed water recovery. She et al [91] demonstrated that a higher water flux would
induce a greater hydrodynamic drag force which enhanced the formation of the fouling layer on the

us
membrane surface. A Similar observation was reported by Yip et al [95] and it was found that the
foulant materials from the feed would accumulate into the membrane porous support layer carried by
the permeation water and caused a severe water flux decline. Chemical-free osmotic backwash was

an
demonstrated to be effective in removing the foulants from the support layer [95].
Recently, Wan et al [96] experimentally studied the PRO fouling of the PES TFC-PRO hollow fiber
membrane by using real seawater brine from a SWRO desalination plant as the draw solution and real
M
wastewater retentate from a wastewater treatment plant as the feed. As shown in Fig. 24, a high power
density of 27.0 W/m2 and 21.1 W/m2 could be achieved at 20 bar using deionized water as the feed, 1 M
NaCl solution and the seawater brine as the draw solutions, respectively. However, the power densities
dramatically dropped to 4.6 W/m2 when the wastewater retentate was used as the feed solution. They
d

found that fouling on the membrane porous substrate induced by the wastewater feed was very fast
te

and severe which could cause a 75-80% reduction in water flux; while the fouling tendency of the
seawater brine on the selective skin was negligible. This is because the membrane selective skin
hindered the penetration of the foulants in the seawater brine into the membrane, and the water
p

permeation flux tended to wash off the foulants on the skin. Clearly, fouling on PRO membranes must
be mitigated in order to sustain high performance of the full-scale PRO processes.
ce

Figure 24
Li et al [97] recently molecularly designed anti-fouling TFC-PRO hollow fiber membranes by
synthesizing a dendritic hyper-branched polyglycerol (HPG) with well-controlled grafting sites, and
Ac

grafting it on PES hollow fiber membrane supports with the aid of polydopamine, as illustrated in Fig. 25.
The grafting route provides a properly-controlled way to form a monolayer of the dendritic polymer on
the substrate. Not only can it effectively cover the substrate surface but also prevent pore blockage.
Anti-fouling assays were then conducted to investigate the fouling resistance of modified membranes
and evaluate their performance in PRO tests. The HPG grafted membranes showed good anti-fouling
effects against Bovine Serum Albumin (BSA) adsorption, E. coli adhesion, and S. aureus attachment, and
exhibited a high flux recovery up to 94% after cleaning and hydraulic pressure impulsion. Later, Cai et al
[98] demonstrated another effective antifouling strategy by grafting zwitterionic copolymers onto the
PRO membrane surfaces. The grafted membranes not only exhibited superior fouling resistance against
BSA adsorption, E. coli adhesion, and S. epidermidis attachment, but also showed substantial improved

22
Page 22 of 42
water flux recovery up to 98% in the PRO tests using real concentrated municipal wastewater. These
works may provide effective means for the molecular design of antifouling PRO membranes.
Figure 25
Generally, three strategies have been identified to mitigate the membrane fouling in PRO: (1) Pre-
treating the feed water to reduce the fouling potential; (2) Manipulating the structural properties of the
membranes to decrease its affinity towards foulants; (3) Cleaning the membranes during operation by

t
backwashing, chemical cleaning, osmotic shock or other methods [89-98]. A preferred antifouling

ip
strategy should effectively mitigate fouling formation without significantly increasing the operating cost
of the process, which makes the development of specific antifouling PRO membranes especially
attractive.

cr
8. Conclusion and Perspectives

us
Pressure retarded osmosis (PRO) for osmotic power generation has been significantly advanced in the
last few years. Several high performance flat-sheet and hollow fiber PRO membranes with appropriate

an
membrane structure, transport properties and robustness have been developed because of the
advances in understanding PRO processes and breakthroughs in new materials and novel fabrication
strategies. At this point of time, the best PRO membranes can withstand a pressure larger than 20 bar
and perform a power density higher than 20 W/m2. However, it should be noted that most reported
M
PRO performance were based on laboratory PRO setups without turbines. In addition, most of them
used synthetic feeds, i.e., clean draw and feed solutions containing only NaCl instead of real brine and
water sources. Furthermore, the actual power output from a PRO plant is not only from membranes’
d

power density, but also dependent on turbine design and efficiency as well as the energy recovery
devices. The real potential of the newly developed PRO membranes for osmotic power generation has
te

yet to be proven with long-term operations in PRO plants.


For real PRO applications, the PRO membranes must be configured into modules. However, so far
p

only limited studies have been reported to investigate what kind of membrane modules could achieve
high efficiency and power output. The conventional module designs for current water treatments show
ce

severe limitations for PRO applications in terms of spacer, internal flow pattern, pressure loss,
membrane area, and membrane deformation [21,53,89]. In addition, modelled results indicate that the
trade-off between power density and specific energy might be an inherent challenge to the real PRO
Ac

plants at module scale [99]. Therefore, comprehensive design criteria are urgently required for PRO
membrane modules such as module configuration, flow pattern and membrane spacer. Pioneered by
Statkraft, more companies around the world are looking into the fabrication of effective PRO
membranes and modules including Hydration Technology Innovations (HTI), Nitto Denko/Hydranautics,
General Electric, Toray Industries and Koch Membrane [25]. Further progresses in the design and
fabrication of effective PRO membrane modules are foreseen.
Another key issue for PRO is the membrane fouling, which is closely linked to (1) quality of the draw
and feed waters, (2) pre-treatments of the water streams, (3) membrane characteristics, and (4)
membrane module design. So far, only few studies have touched this interesting area due to the lack of
commercially available PRO membranes and membrane modules. Further investigations using large
membrane modules are needed for better understanding of fouling and cleaning in PRO processes

23
Page 23 of 42
because the operating conditions in PRO plants are different from those of laboratory scale PRO cells.
More R & D efforts on membrane fouling, antifouling membranes and membrane cleaning are expected
for PRO in the near future.
Recently, other interesting PRO schemes using alternative hypersaline waters, such as the brine from
RO desalination plants, are emerging after the traditional pairing of river water and sea water [9-
11,53,100]. Due to their much higher salinities, the membrane power density is projected to be

t
improved significantly [101,102]. However, lower efficiency and higher operating hydraulic pressures are

ip
expected because of severe concentration polarization and much larger osmotic pressure gradients.
Therefore, PRO membranes with a smaller structural parameter and higher mechanical robustness must
be developed. In addition, the development of sustainable processes by integrating PRO power

cr
generation and water purification has gained attention recently. For example, a successful integration of
PRO and RO plants could utilize the high salinity RO retentate as the saltwater for PRO power generation

us
while the diluted saltwater can be used as the feed for RO desalination [101-104]. Not only can this
combination lower the overall energy consumption for SWRO, but also solve the disposal problem of RO
retentate.

an
Frankly, the PRO technology for osmotic power generation is still in an infancy stage and it takes time
for industries to integrate PRO with RO or other processes; however, technology evolutions on both
membranes and system optimization have been taking place continuously to bring PRO towards
M
commercialization. Driven by the huge economical values of producing PRO membranes and enormous
environmental benefits of integrating PRO processes with RO, more membrane manufacturers are
looking into these new technologies. Encouragingly, breakthroughs on high performance PRO
membranes have been demonstrated recently by academic membrane scientists. Future works should
d

aim at cooperation between academic institutes and membrane manufactures to further advance
te

module fabrication and anti-fouling membranes so that high efficient PRO plants can be developed as
part of conventional desalination processes in the near future.
p

Acknowledgments
ce

This research grant is supported by the Singapore National Research Foundation, Prime Minister’s
Office, Singapore under its Environmental & Water Technologies Strategic Research Programme and
administered by the Environment & Water Industry Programme Office (EWI) of the PUB under the
Ac

project entitled “Membrane development for osmotic power generation, Part 1. Materials development
and membrane fabrication” (1102-IRIS-11-01) and NUS grant number of R-279-000-381-279.

24
Page 24 of 42
References

[1] Anonymous. International Energy Outlook 2013. US Energy Information


Administration, www.eia.gov/oiaf/ieo/index.html 2013. 62 pp.
[2] Evans A, Strezov V, Evans TJ. Assessment of sustainability indicators for

t
ip
renewable energy technologies. Renew Sust Energ Rev 2009;13:1082-1088.
[3] Pattle RE. Production of electric power by mixing fresh and salt water in the

cr
hydroelectric pile. Nature 1954;174:660-660.
[4] Loeb S, Norman RS. Osmotic power plants. Science 1975;189:654-655.

us
[5] Yip NY, Elimelech M. Comparison of energy efficiency and power density in
pressure retarded osmosis and reverse electrodialysis. Environ Sci Technol

an
2014;48:11002-11012.
[6] Gerstandt K, Peinemann KV, Skilhagen SE, Thorsen T, Holt T. Membrane
processes in energy supply for an osmotic power plant. Desalination 2008;224:64-
M
70.
[7] Loeb S. Production of energy from concentrated brines by pressure-retarded
d

osmosis: I. Preliminary technical and economic correlations. J Membr Sci


1976;1:49-63.
te

[8] Thorsen T, Holt T. The potential for power production from salinity gradients by
p

pressure retarded osmosis. J Membr Sci 2009;335:103-110.


[9] Post JW, Veerman J, Hamelers HVM, Euverink GJW, Metzb SJ, Nymeijer K,
ce

Buisman CJN. Salinity-gradient power: Evaluation of pressure-retarded osmosis


and reverse electrodialysis. J Membr Sci 2007;288:218-230.
Ac

[10] Achilli A, Childress AE. Pressure retarded osmosis: from the vision of Sidney Loeb
to the first prototype installation—review. Desalination 2010;261:205-211.
[11] Chung TS, Li X, Ong RC, Ge QC, Wang HL, Han G. Emerging forward osmosis
(FO) technologies and challenges ahead for clean water and clean energy
applications. Curr Opin Chem Eng 2012;1:246-257.
[12] Logan BE, Elimelech M. Membrane-based processes for sustainable power
generation using water. Nature 2012;488:313-319.
[13] Norman RS. Water desalination: A source of energy. Science 1974;186:350-352.

25
Page 25 of 42
[14] Loeb S, Hessen FV, Shahaf D. Production of energy from concentrated brines by
pressure retarded osmosis. II. Experimental results and projected energy costs. J
Membr Sci 1976;1:249-269.
[15] Mehta GD, Loeb S. Performance of permasep B-9 and B-10 membranes in various
osmotic regions and at high osmotic pressures. J Membr Sci 1979;4:335-349.

t
ip
[16] Mehta GD, Loeb S. Internal polarization in the porous substructure of a
semipermeable membrane under pressure-retarded osmosis. J Membr Sci

cr
1978;4:261-265.
[17] Lee KL, Baker RW, Lonsdale HK. Membranes for power generation by pressure-

us
950 retarded osmosis. J Membr Sci 1981;8:141-171.
[18] Anonymous Statkraft Website http://www.statkraft.com 2014.

an
[19] Skilhagen SE, Dugstad JE, Aaberg RJ. Osmotic power—power production based on
the osmotic pressure difference between waters with varying salt gradients.
Desalination 2008;220:476-482.
M
[20] Saito K, Irie M, Zaitsu S, Sakai H, Hayashi H, Tanioka A. Power generation with
salinity gradient by pressure retarded osmosis using concentrated brine from SWRO
d

system and treated sewage as pure water. Desalination Water Treat 2012;41:114-
121.
te

[21] Skilhagen SE. Osmotic power- a new, renewable energy source. Desalination Water
p

Treat 2010;15:271-278.
[22] Alsvik IL, Hägg MB. Pressure retarded osmosis and forward osmosis membranes:
ce

materials and methods. Polymers 2013;5:303-327.


[23] Zhao S, Zou L, Tang CY, Mulcahy D. Recent developments in forward osmosis:
Ac

Opportunities and challenges. J Membr Sci 2012;396:1-21.


[24] Achilli A, Cath TY, Childress AE. Power generation with pressure retarded
osmosis: An experimental and theoretical investigation. J Membr Sci 2009;343:42-
52.
[25] Helfer F, Lemckert C, Anissimov YG. Osmotic power with pressure retarded
osmosis: theory, performance and trends-A review. J Membr Sci 2014;453:337-358.

26
Page 26 of 42
[26] Klaysom C, Cath TY, Depuydt T, Vankelecom IFJ. Forward and pressure retarded
osmosis: potential solutions for global challenges in energy and water supply. Chem
Soc Rev 2013;42:6959-6989.
[27] Han G, Ge Q, Chung TS. Conceptual demonstration of novel closed-loop pressure
retarded osmosis process for sustainable osmotic energy generation. Appl Energy

t
ip
2014;132:383-393.
[28] Cath TY, Childress AE, Elimelech M. Forward osmosis: principles, applications,

cr
and recent developments. J Membr Sci 2006;281:70-87.
[29] Thorsen T, Holt T. Hydrophile semipermeable membrane. WO Pat 047733 A1

us
2003.
[30] Thorsen T, Holt T. Semi permeable membrane for use (56) in osmosis, and method

an
and plant for providing elevated pressure by osmosis to create power. US Pat,
7,566,402 B2 2009.
[31] Dinger F, Trondle T, Platt U. Optimization of the energy output of osmotic power
M
plants. J Renew Energy 2013;7:496768/1-7.
[32] Loeb S, Honda T, Reali M. Comparative mechanical efficiency of several plant
d

configurations using a pressure-retarded osmosis energy converter. J Membr Sci


1990;51:323-335.
te

[33] Yip NY, Elimelech M. Performance limiting effects in power generation from
p

salinity gradients by pressure retarded osmosis. Environ Sci Technol


2011;45:10273-10282.
ce

[34] Loeb S, Titelman L, Korngold E, Freiman J. Effect of porous support fabric on


osmosis through a Loeb-Sourirajan type asymmetric membrane. J Membr Sci
Ac

1997;129:243-249.
[35] Hoover LA, Schiffman JD, Elimelech M. Nanofibers in thin-film composite
membrane support layers: Enabling expanded application of forward and pressure
retarded osmosis. Desalination 2013;308:73-81.
[36] Zelman A. Membrane permeability: Generalization of the reflection coefficient
method of describing volume and solute flows. Biophys J 1972;12:414-419.

27
Page 27 of 42
[37] Yip NY, Tiraferri A, Phillip WA, Schiffman JD, Hoover LA, Kim YC, Elimelech
M. Thin-film composite pressure retarded osmosis membranes for sustainable
power generation from salinity gradients. Environ Sci Technol 2011;45:4360-4369.
[38] Han G, Zhang S, Li X, Chung TS. High performance thin film composite pressure
retarded osmosis (PRO) membranes for renewable salinity-gradient energy

t
ip
generation. J Membr Sci 2013;440:108-121.
[39] Zhang S, Chung TS. Minimizing the instant and accumulative effects of salt

cr
permeability to sustain ultrahigh osmotic power density. Environ Sci Technol
2013;47:10085-10092.

us
[40] Geise GM, Lee HS, Miller DJ, Freeman BD, McGrath JE, Paul DR. Water
purification by membranes: the role of polymer science. J Polym Sci Part B Polym

an
Phys 2010;48:1685-1718.
[41] Zhang S, Zhang RW, Jean YC, Paul DR, Chung TS. Cellulose esters for forward
osmosis: characterization of water and salt transport properties and free volume.
M
Polymer 2012;53:2664-2672.
[42] Widjojo N, Chung TS, Weber M, Maletzko C, Warzelhan V. The role of
d

sulphonated polymer and macrovoid-free structure in the support layer for thin-film
composite (TFC) forward osmosis (FO) membranes. J Membr Sci 2011;383:214-
te

223.
p

[43] Wang KY, Chung TS, Amy G. Developing Thin-film-composite forward osmosis
membranes based on the PES/SPSf substrate through interfacial polymerization.
ce

AIChE J 2012;58:770-781.
[44] Li X, Zhang S, Fu FJ, Chung TS. Deformation and reinforcement of thin-film
Ac

composite (TFC) polyamide-imide (PAI) membranes for osmotic power generation.


J Membr Sci 2013;434:204-217.
[45] Zhang S, Fu FJ, Chung TS. Substrate modifications and alcohol treatment on thin
film composite, membranes for osmotic power. Chem Eng Sci 2013;87:40-50.
[46] She Q, Jin X, Tang CY. Osmotic power production from salinity gradient resource
by pressure retarded osmosis: effects of operating conditions and reverse solute
diffusion. J Membr Sci 2012;401-402:262-273.

28
Page 28 of 42
[47] Sivertsen E, Holt T, Thelin W, Brekke G. Modelling mass transport in hollow fibre
membranes used for pressure retarded osmosis. J Membr Sci 2012;417-418:69-79.
[48] Banchik LD, Sharqawy MH, Lienhard V JH. Limits of power production due to
finite membrane area in pressure retarded osmosis. J Membr Sci 2014;468:81-89.
[49] Loeb S, Mehta GD. A two-coefficient water transport equation for pressure-

t
ip
retarded osmosis. J Membr Sci 1978;4:351-362.
[50] Jellinek HH, Masuda H. Osmo-power. Theory and performance of an osmo-power

cr
pilot plant. Ocean Eng 1981;8:103-128.
[51] Mehta GD. Further results on the performance of present-day osmotic membranes

us
in various osmotic regions. J Membr Sci 1982;10:3-19.
[52] McCutcheon JR, Elimelech M. Influence of membrane support layer

an
hydrophobicity on water flux in osmotically driven membrane processes. J Membr
Sci 2008;318:458-466.
[53] Kim YC, Elimelech M. Potential of osmotic power generation by pressure retarded
M
osmosis using seawater as feed solution: Analysis and experiments. J Membr Sci
2013;429:330-337.
d

[54] Xu Y, Peng XY, Tang CY, Fu QSA, Nie SZ. Effect of draw solution concentration
and operating conditions on forward osmosis and pressure retarded osmosis
te

performance in a spiral wound module. J Membr Sci 2010;348:298-309.


p

[55] Han G, Zhang S, Li X, Widjojo N, Chung TS. Thin film composite forward osmosis
membranes based on polydopamine modified polysulfone substrates with
ce

enhancements in both water flux and salt rejection. Chem Eng Sci 2012;80:219-231.
[56] Han G, Chung TS, Toriida M, Tamai S. Thin-film composite forward osmosis
Ac

membranes with novel hydrophilic supports for desalination. J Membr Sci


2012;423-424:543-555.
[57] Cadotte JE. Reverse osmosis membrane. US 4,039,440, 1977.
[58] Cadotte JE, Petersen RJ, Larson RE, Erickson EE. A new thin-film composite
seawater reverse osmosis membrane. Desalination 1980;32:25-31.
[59] Singh PS, Joshi SV, Trivedi JJ, Devmurari CV, Rao AP, Ghosh PK. Probing the
structural variations of thin film composite RO membranes obtained by coating

29
Page 29 of 42
polyamide over polysulfone membranes of different pore dimensions. J Membr Sci
2006;278:19-25.
[60] Ghosh AK, Hoek EMV. Impacts of support membrane structure and chemistry on
polyamide-polysulfone interfacial composite membranes. J Membr Sci
2009;336:140-148.

t
ip
[61] Kong CL, Kanezashi M, Yamomoto T, Shintani T, Tsuru T. Controlled synthesis of
high performance polyamide membrane with thin dense layer for water desalination.

cr
J Membr Sci 2010;362:76-80.
[62] Li X, Wang KY, Helmer BJ, Chung TS. Thin-film composite membranes and

us
formation mechanism of thin-film layer on hydrophilic cellulose acetate propionate
substrates for forward osmosis processes. Ind Eng Chem Res 2012;51:10039-10050.

an
[63] Li X, Chung TS. Effects of free volume in thin-film composite membranes on
osmotic power generation. AIChE J 2013;59:4749-4761.
[64] Cui Y, Liu XY, Chung TS. Enhanced osmotic energy generation from salinity
M
gradients by modifying thin film composite membranes. Chem Eng J
2014;242:195-203.
d

[65] Zuo J, Wang Y, Sun SP, Chung TS. Molecular design of thin film composite (TFC)
hollow fiber membranes for isopropanol dehydration via pervaporation. J Membr
te

Sci 2012;405-406:123-133.
p

[66] Tin PS, Chung TS, Hill AJ. Advanced fabrication of carbon molecular sieve
membranes by nonsolvent pretreatment of precursor polymers. Ind Eng Chem Res
ce

2004;43:6476-6483.
[67] Shao L, Chung TS, Goh SH, Pramoda KP. Transport properties of cross-linked
Ac

polyimide membranes induced by different generations of diaminobutane (DAB)


dendrimers. J Membr Sci 2004;238:153-163.
[68] Mansourpanah Y, Madaeni SS, Rahimpour A. Fabrication and development of
interfacial polymerized thin-film composite nanofiltration membrane using different
surfactants in organic phase; study of morphology and performance. J Membr Sci
2009;343:219-228.

30
Page 30 of 42
[69] Kim IC, Jeong BR, Kim SJ, Lee KH. Preparation of high flux thin film composite
polyamide membrane: The effect of alkyl phosphate additives during interfacial
polymerization. Desalination 2013;308:111-114.
[70] Ghosh AK, Jeong BH, Huang XF, Hoek EMV. Impacts of reaction and curing
conditions on polyamide composite reverse osmosis membrane properties. J Membr

t
ip
Sci 2008;311:34-45.
[71] Song X, Liu Z, Sun DD. Energy recovery from concentrated seawater brine by thin-

cr
film nanofiber composite pressure retarded osmosis membranes with high power
density. Energy Environ Sci 2013;6:1199-1210.

us
[72] Bui N, McCutcheon JR. Hydrophilic Nanofibers as New Supports for Thin Film
Composite Membranes for Engineered Osmosis. Environ Sci Technol

an
2013;47:1761-1769.
[73] Reneker DH, Chun I. Nanometre diameter fibres of polymer, produced by
electrospinning. Nanotechnology 1996;7:216.
M
[74] Farr I, Bharwada U, Gillinkala T. Design and performance of HIT’s thin film
composite membrane for forward osmosis and pressure retarded osmosis
d

applications. Procedia Eng 2012;44:1271-1.


[75] Straub AP, Yip NY, Elimelech M. Raising the Bar: Increased Hydraulic Pressure
te

Allows Unprecedented High Power Densities in Pressure-Retarded Osmosis.


p

Environ Sci Technol Lett 2014;1:55-59.


[76] Kim YC, Elimelech M. Adverse impact of feed channel spacers on the performance
ce

of pressure retarded osmosis. Environ Sci Technol 2012;46:4673-4681.


[77] She Q, Hou D, Liu J, Tan KH, Tang CY. Effect of feed spacer induced membrane
Ac

deformation on the performance of pressure retarded osmosis (PRO): Implications


for PRO process operation. J Membr Sci 2013;445:170-182.
[78] Peng N, Widjojo N, Sukitpaneenit P, Teoh MM, Lipscomb GG, Chung TS, Lai JY.
Evolution of polymeric hollow fibers as sustainable technologies: Past, present, and
future. Prog Polym Sci 2012;37:1401-1424.
[79] Sivertsen E, Holt T, Thelin W, Brekke G. Pressure retarded osmosis efficiency for
different hollow fibre membrane module flow configurations. Desalination
2013;312:107-123.

31
Page 31 of 42
[80] Fu FJ, Zhang S, Sun SP, Wang KY, Chung TS. POSS-containing delamination-free
dual-layer hollow fiber membranes for forward osmosis and osmotic power
generation. J Membr Sci 2013;443:144-155.
[81] Fu FJ, Sun SP, Zhang S, Chung TS. Pressure retarded osmosis dual-layer hollow
fiber membranes developed by co-casting method and ammonium persulfate (APS)

t
ip
treatment. J Membr Sci 2014;469:488-498.
[82] Han G, Wang P, Chung TS. Highly robust thin-film composite pressure retarded

cr
osmosis (PRO) hollow fiber membranes with high power densities for renewable
salinity-1122 gradient energy generation. Environ Sci Technol 2013;47:8070-8077.

us
[83] Han G, Chung TS. Robust and high performance pressure retarded osmosis hollow
fiber membranes for osmotic power generation. AIChE J 2014;3:1107-1119.

an
[84] Chou S, Wang R, Shi L, She Q, Tang C, Fane AG. Thin-film composite hollow
fiber membranes for pressure retarded osmosis (PRO) process with high power
density. J Membr Sci 2012;389:25-33.
M
[85] Chou S, Wang R, Fane AG. Robust and High performance hollow fiber membranes
for energy harvesting from salinity gradients by pressure retarded osmosis. J
d

Membr Sci 2013;448:44-54.


[86] Li X, Chung TS. Thin-film composite P84 co-polyimide hollow fiber membranes
te

for osmotic power generation. Appl Energy 2014;114:600-610.


p

[87] Zhang S, Sukitpaneenit P, Chung TS. Design of robust hollow fiber membranes
with high power density for osmotic energy production. Chem Eng J 2014;241:457-
ce

465.
[88] Sun SP, Chung TS. Outer-Selective Pressure-Retarded Osmosis Hollow Fiber
Ac

Membranes from Vacuum-Assisted Interfacial Polymerization for Osmotic Power


Generation. Environ Sci Technol 2013;47:13167-13174.
[89] Kleverud J, Skilhagen SE, Brekke G. Experiences with the Tofte prototype plant.
Proceedings of the 3rd Osmosis Membrane Summit, Barcelona, Statkraf, 2012. 28
pp. http://osmosis-
summit.event123.no/pop.cfm?FuseAction=Doc&pAction=View&pDocumentId=37
222

32
Page 32 of 42
[90] Zhang M, Hou D, She Q, Tang CY. Gypsum scaling in pressure retarded osmosis:
Experiments, mechanisms and implications. Water Res 2014;48:387-395.
[91] She Q, Wong YKW, Zhao S, Tang CY. Organic fouling in pressure retarded
osmosis: Experiments, mechanisms and implications. J Membr Sci 2013;428:181-
189.

t
ip
[92] Chen SC, Fu XZ, Chung TS. Fouling behaviors of polybenzimidazole (PBI)-1146
polyhedral oligomeric silsesquioxane (POSS)/polyacrylonitrile (PAN) hollow fiber

cr
membranes for engineering osmosis processes. Desalination 2014;335:17-26.
[93] Chen SC, Wan CF, Chung TS. Enhanced fouling by inorganic and organic foulants

us
on pressure retarded osmosis (PRO) hollow fiber membranes under high pressures.
J Membr Sci 2015;479:190-203.

an
[94] Thelin WR, Sivertsen E, Holt T, Brekke G. Natural organic matter fouling in
pressure retarded osmosis. J Membr Sci 2013;438:46-56.
[95] Yip NY, Elimelech M. Influence of natural organic matter fouling and osmotic
M
backwash on pressure retarded osmosis energy production from natural salinity
gradients. Environ Sci Technol 2013;47:12607-12616.
d

[96] Wan CF, Chung TS. Osmotic power generation by pressure retarded osmosis using
seawater brine as the draw solution and wastewater retentate as the feed. J Membr
te

Sci 2015;479:148-158.
p

[97] Li X, Cai T, Chung TS. Anti-fouling behavior of hyper-branched polyglycerol


grafted polyethersulfone hollow fiber membranes for osmotic power generation.
ce

Environ Sci Technol 2014;48:9898-9907.


[98] Cai T, Li X, Wan CF, Chung TS. Zwitterionic polymers grafted poly(ether sulfone)
Ac

hollow fiber membranes and their antifouling behaviors for osmotic power
generation. Environ Sci Technol 2015;submitted.
[99] Straub AP, Lin S, Elimelech M. Module-scale analysis of pressure retarded
osmosis: performance limitations and implications for full-Scale operation. Environ
Sci Technol 2014;48:12435-12444.
[100] Lin S, Straub AP, Elimelech M. Thermodynamic limits of extractable energy by
pressure retarded osmosis. Energy Environ Sci 2014;7:2706-2714.

33
Page 33 of 42
[101] Efraty A. Closed Circuit PRO Series No 5: clean energy generation from seawater
and its concentrates by CC-PRO without need of energy recovery. Desalination
Water Treat 2015;DOI:10.1080/19443994.2015.1017323.
[102] Chung TS, Zhang S, Wang KY, Su JC, Ling MM. Forward osmosis processes:
yesterday, today and tomorrow. Desalination 2012;287:78-81.

t
ip
[103] Achilli A, Prante JL, Hancock NT, Maxwell EB, Childress AE. Experimental
results from RO-PRO: A next generation system for low-energy desalination.

cr
Environ Sci Technol 2014;48:6437-6443.
[104] Kim J, Park M, Snyder SA, Kim JH. Reverse osmosis (RO) and pressure retarded

us
osmosis (PRO) hybrid processes: Model-based scenario study. Desalination
2013;322:121-130.

an
M
[1] Maximum energy that could be theoretically extracted from the mixing of freshwater
with salty water from different sources. [25].

[2] Schematic representations of the osmotic processes.


d

[3] Schematic diagram of a typical PRO osmotic power plant with continuous and steady state
te

flow.

[4] Schematic of the salt concentration and osmotic pressure profiles across an
p

asymmetric composite membrane operated in PRO. Cs and Cf are the salt


concentrations of the bulk fresh water and salty water solutions, respectively. C1 is the
ce

salt concentration at the interface between the salty water and the membrane selective
layer. C2 is the salt concentration at the interface between the membrane support
and the active layer. C3 is the salt concentration at the interface of the fresh water
Ac

and the membrane substrate. ∆πbulk is the osmotic pressure different between the
bulk fresh water and salty water. ∆πm is the osmotic pressure different across the
membrane. ∆πeff is the effective osmotic pressure gradient across the membrane
selective layer.
[5] Representative plots of (a) water flux Jw and (b) power density W as a function of
applied hydraulic pressure difference ∆P. The ideal Jw and W without any detrimental
effect are indicated by the solid gray line and calculated using Eq. (7) and Eq. (8),
respectively. The solid dark line represents the actual water flux (Jact) and actual
power density (Wact) calculated using Eq. (16) and Eq. (17), respectively. The dashed
lines indicate the Jw and W when each of the detrimental effects are absent

34
Page 34 of 42
(ICP=internal concentration polarization, ECP=external concentration polarization
and JSflux). Jw and W were calculated using πs = 26.14 bar, πf = 0.789 bar, A = 4.0 L
m-2 h-1 bar-1, B = 0.85 L m-2 h-1, S = 350 µm, and k = 38.5 µm/s. [33], Copyright
2011. Reproduced with permission from the American Chemical Society.
[6] Iso-watt diagrams constructed for constant structure parameter (a) S=0.5 mm, (b)
S=3.0 mm, and constant structure parameter diagrams constructed for a given

t
specific power (c) W=1.5 W/m2, (d) W=4.5 W/m2. Calculations are based on an

ip
average salt water concentration equal to 28 giL, film thicknesses ds=d1=50 J..Lm
and temperature 20 °C. [47], Copyright 2012. Reproduced with permission from

cr
Elsevier Ltd.

[7] Power densities W (W/m2) obtained in PRO using commercial available RO/NF

us
membranes. The power density is derived from the reported osmotic pressure,
hydraulic pressure, and from water fluxes (W1-W2: Loeb et al. [7,14]; W12: Jellinek
and Masuda [50]) or from permeation coefficients (W3-W7: Mehta and Loeb [15,16];

an
W8: Loeb and Mehta [49]; W9-W11: Mehta [51]).

[8] SEM images of the CTA-FO membranes from Hydration Technology Innovations
M
(HTI): (a) FO-1, (b) FO-2. [11], [52], Copyright 2012, 2008, respectively..
Reproduced with permission from Elsevier Ltd.

[9] Power densities W (W/m2) obtained with PRO using CTA-FO membranes. The power
d

density is derived from reported osmotic pressure, hydraulic pressure, and water
te

fluxes [6,19,23-26,46].

[10] SEM images of the polyamide-imide (PAI) membrane before and after being
p

compacted in PRO at 14 bar for 250 min. [44], Copyright 2013. Reproduced with
permission from Elsevier Ltd.
ce

[11] (a, b) FESEM images of the Matrimid® TFC-PRO membrane. [38], Copyright 2013.
Reproduced with permission from Elsevier Ltd; (c) the nanofiber membrane substrate
Ac

(top surface) with its nanofiber diameter distribution (inserted plot); and (d) the cross
section of the TFC-PRO membrane. [71], Copyright 2013. Reproduced with
permission from the Royal Society of Chemistry.

[12] (a) SEM images of the Matrimid® membrane substrate, (b) the TFC-PRO membrane,
and (c) the power density as a function of post-treatments using 1 M NaCl as the draw
solution and deionized water as the feed. [64], Copyright 2014. Reproduced with
permission from Elsevier Ltd.

[13] (a) Membrane deformation under the PRO operation using different feed spacers; (b)
membrane separation parameters (B/A) (b) and (c) power density (W) as a function

35
Page 35 of 42
applied hydraulic pressure difference (∆P) for different spacers [77], Copyright 2013.
Reproduced with permission from Elsevier Ltd.

[14] Cross-section morphology and PRO performances of the PBI/POSS–PAN/PVP hollow


fiber membrane with a POSS loading of 0.5 wt%. (a) Cross-section of the fiber; (b)
Cross-section of interfaces; (c) Cross-section of the outer selective layer; (d) Water
permeation flux; (d) power density in PRO. Draw solution: 1M NaCl; feed solution:

t
0.01 M NaCl [80], Copyright 2013. Reproduced with permission from Elsevier Ltd.

ip
[15] SEM images of the outer surface and cross-section morphology of the Matrimid®
hollow fiber membrane substrates spun under different conditions, [83], Copyright

cr
2014. Reproduced with permission from the American Institute of Chemical
Engineers.

us
[16] The effects of pre-stabilization on the transport properties of the TFC-PRO hollow
fiber membranes with different substrates shown in Figure 15 (the stabilization
pressures were 16, 15, 15, and 6 bar, respectively), [83], Copyright 2014.

an
Reproduced with permission from the American Institute of Chemical Engineers.

[17] Schematic of the membrane properties variations of the inner-selective TFC-PRO


hollow fiber membranes during PRO operations [82], Copyright 2013. Reproduced
M
with permission from the American Chemical Society.

[18] SEM images of the TFC-PRO membranes with a polyetherimide (PEI) hollow fiber
d

substrate (a-c) [85], Copyright 2013. Reproduced with permission from Elsevier Ltd,
and a P84 co-polyimide hollow fiber substrate (d-f). [86], Copyright 2014.
te

Reproduced with permission from Elsevier Ltd.

[19] SEM images of the cross-section and surface morphologies of the poly(ether sulfone)
p

(PES) hollow fiber substrate. [39], Copyright 2013. Reproduced with permission from
ce

the American Chemical Society; [87] Copyright 2014. Reproduced with permission
from Elsevier Ltd.

[20] The (a) power density and (b) specific reverse salt flux (Js/Jw) of the stabilized
Ac

poly(ether sulfone) (PES) TFC- PRO hollow fiber membranes in the PRO process for
osmotic power generation. The draw solution is 1 M NaCl, and feed solution is
deionized water. [87], Copyright 2014. Reproduced with permission from Elsevier Ltd.

[21] Structures of the reported polymer materials for the fabrication of high performance
TFC-PRO hollow fiber membranes.

[22] Schematic illustration of the desirable highly asymmetric hollow fiber substrate for
inner-selective TFC- PRO hollow fiber membranes for osmotic power generation. [87],
Copyright 2014. Reproduced with permission from Elsevier Ltd.

36
Page 36 of 42
[23] (a) Cross-section morphology of the Matrimid® hollow fiber support, (b) outer
surface morphology of the TFC-PRO membrane, (c) water permeability (A) and salt
permeability (B) of the TFC-PRO hollow fiber membranes tested by RO experiments
with a constant pressure of 1 bar and a NaCl concentration of 200 ppm, (d) power
density (W) of the TFC-PRO hollow fiber membranes. The draw solution is 1 M NaCI
and the feed solution is deionized water. [88], Copyright 2013. Reproduced with
permission from the American Chemical Society.

t
ip
[24] (a) Water flux and (b) power density of the poly(ether sulfone) (PES) TFC-PRO
hollow fiber membrane as functions of transmembrane pressure difference (∆P).

cr
Baseline: 1M NaCl solution as the draw solution (DS) and DI water as the feed
solution (FS), (A) 0.81M NaCl solution as DS and DI water as FS, (B) SWRO brine
(SWBr) as DS and DI water as FS, (C) 0.81M NaCl solution as DS and the real

us
wastewater retentate (WWRe) brine as FS, and (D) SWBr as DS and WWRe as FS.
[96], Copyright 2015. Reproduced with permission from Elsevier Ltd.

[25] Anti-fouling poly(ether sulfone) (PES) TFC·PRO hollow fiber membranes for

an
osmotic power generation modified by hyperbranched polyglycerol. [97], Copyright
2014. Reproduced with permission from the American Chemical Society.
M
Table 1 Transport properties of the commercial CTA-FO membranes from HTI
d

Water permeability, A Salt permeability, B S (µm) Reference


-9 -1 -1 -7 -1
(10 m s kPa ) (10 m s )
te

1.87 1.11 678 [24]


p

1.02 0.77 590 [46]


ce

1.00 0.89 505 [53]


Ac

1.83 1.20 790 [72]

3.42 7.30 689 [76]

2.08 1.76 480 [77]

Membrane: Commercial HTI CTA-FO membrane

37
Page 37 of 42
t
ip
cr
us
an
M
Table 2 Summary of the PRO performance of the reported TFC-PRO flat-sheet
membranes
Membrane Draw solution Feed ∆P W Reference
d

2
(bar) (W/m )
te

TFC-1 Synthetic RO brine Deionized water 9.0 6.0 [63]


(1 M NaCl)
p

TFC-2 Synthetic RO brine Deionized water 9.0 5.3 [63]


(1 M NaCl)
ce

TFC-3 Synthetic RO brine Deionized water 15.0 12.0 [38]


(1 M NaCl)
Ac

Synthetic RO brine Synthetic river 15.0 10.0 [38]


(1 M NaCl) water (10 mM
NaCl)
Synthetic RO brine Synthetic 15 9.2 [38]
(1 M NaCl) wastewater brine
(40 mM NaCl)
Synthetic seawater Deionized water 13.0 9.0 [38]
(0.59 M NaCl)
TFC-4 Synthetic RO brine Deionized water 22.0 18.1 [64]
(1 M NaCl)

38
Page 38 of 42
TFC-5 Synthetic RO brine Deionized water 22.0 16.9 [64]
(1 M NaCl)
TFC-6 Synthetic RO brine Deionized water 22.0 15.8 [64]
(1 M NaCl)
TFC-7 Synthetic RO brine Deionized water 22.0 8.6 [64]
(1 M NaCl)

t
ip
TFC-8 Synthetic seawater Synthetic river 15.2 21.3 [71]
brine (1.06 M water (0.9 mM
NaCl) NaCl)

cr
Synthetic seawater Synthetic brackish 15.2 15.2 [71]
brine (1.06 M water

us
NaCl) (80 mM NaCl)
TFC-9 Synthetic seawater Deionized water 11.5 8.0 [72]
(0.5 M NaCl)

an
TFC-10 Synthetic RO brine Deionized water 20.7 14.1 [75]
(1 M NaCl)
Synthetic seawater Deionized water 13.8 7.5 [75]
M
(0.6 M NaCl)
d

Table 3 Transport properties of the reported TFC-PRO flat-sheet membranes and the HTI
CTA membrane
te

-9 -1 -1 -7 -1
Membrane A (10 m s kPa ) B (10 m s ) S (µm) Reference
p

TFC-1 2.78 0.50 - [63]


ce

TFC-2 2.22 0.30 - [63]

TFC-3 14.72 5.55 600 [38]


Ac

TFC-4 2.77 - - [64]

TFC-5 2.20 - - [64]

TFC-6 1.52 - - [64]

TFC-7 1.19 - - [64]

TFC-8 11.39 4.83 150 [71]

39
Page 39 of 42
TFC-9 7.86 1.22 273 [72]

HTI-TFC 6.92 1.08 564 [75]

HTI-CTA 1.83 1.20 790 [72]

t
ip
cr
us
an
M
d

Table 4 Transport properties and the burst pressure and maximum power density of the
te

reported TFC-PRO hollow fiber membranes


-9 -7
Membrane A (10 m B (10 Wmax Burst S Reference
-1 -1 -1 2 pressure (µm)
p

s kPa ) ms ) (W/m )
(bar)
ce

TFC-1 11.94 1.30 16.5 >16 640 [83]

TFC-2 10.00 2.03 14.0 >16 640 [83]


Ac

TFC-3 9.22 0.39 10.6 9.5 460 [84]

TFC-4 4.22 0.67 20.9 16.5 610 [85]

TFC-5 2.52 0.24 12.0 24 685 [86]

TFC-6 9.17 0.86 24.3 21 450 [87]

TFC-7 3.89 0.33 13.0 17.5 510 [87]

40
Page 40 of 42
TFC-8 2.50 1.11 7.3 20 540 [87]

t
ip
cr
us
an
M
d
te

Table 5 Characteristics of the hollow fiber substrates for the reported TFC-PRO hollow
fiber membranes
p

Hollow Material Spinning Mean MWCO Porosity Fiber Reference


fiber method pore (kDa) (%) OD/ID
ce

substrate diameter (µm)


(nm)
®
HF-1 Matrimid Dual-bath 13.7 146.4 70.3 820/520 [83]
Ac

coagulation
HF-2 Polyethersulfone Normal dry- 9.1 45 80 1380/980 [84]
jet wet
HF-3 Polyetherimide Normal dry- - 321 72.0 1260/975 [85]
jet wet
HF-4 P84 co- Dual-layer 3.51 26.1 - 1011/630 [86]
polyimide co-extrusion
HF-5 Polyethersulfone Dual-layer 8.7 - 75.3 973/562 [87]
co-extrusion

41
Page 41 of 42
Ac
ce
pte
d
M
an
us
cr
ip
t

42
Page 42 of 42

Das könnte Ihnen auch gefallen