You are on page 1of 9

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Review

Cite This: Environ. Sci. Technol. Lett. 2018, 5, 467−475 pubs.acs.org/journal/estlcu

High-Pressure Reverse Osmosis for Energy-Efficient Hypersaline


Brine Desalination: Current Status, Design Considerations, and
Research Needs
Douglas M. Davenport, Akshay Deshmukh, Jay R. Werber, and Menachem Elimelech*
Department of Chemical and Environmental Engineering, Yale University, New Haven, Connecticut 06520-8286, United States
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Water scarcity, expected to become more widespread in the


coming years, demands renewed attention to freshwater protection and
management. Critical to this effort are the minimization of freshwater
withdrawals and elimination of wastewater discharge, both of which can be
achieved via zero liquid discharge (ZLD), an aggressive wastewater management
Downloaded via 95.154.221.91 on October 2, 2021 at 12:30:28 (UTC).

approach. Because of the high energetic cost of thermal desalination, ZLD is


particularly challenging for high-salinity wastewaters. In this review, we discuss
the potential of high-pressure reverse osmosis (HPRO) (i.e., reverse osmosis
operated at a hydraulic pressure greater than ∼100 bar) to efficiently desalinate
hypersaline brines. We first discuss the inherent energy efficiency of membrane processes compared to that of conventional
thermal processes for brine desalination. We then highlight the opportunity of HPRO to reduce energy requirements for
desalination of key high-salinity industrial wastewaters. The current state of membrane materials and processes for hypersaline
brine desalination is also discussed, emphasizing several process design considerations unique to HPRO. Lastly, we discuss the
most pressing research needs for the development of HPRO, notably the development of membranes and modules suitable for
high pressures as well as fundamental studies of compaction and transport under HPRO conditions.

■ INTRODUCTION
Recent estimates suggest 1.2 billion people live in areas of
flotation, chemical precipitation, or advanced oxidation to
remove oil and grease and reduce organic content.9 Produced
physical water scarcity.1 Effective water resource management water treated in this fashion is often discharged to the ocean at
is critical to ensure the widespread availability of freshwater in offshore locations, albeit with some environmental concerns.10
the coming decades.2 Two specific concerns are the treatment At inland sites, treated produced water can be reused in
and disposal of wastewater and the utilization of previously hydraulic fracturing fluid.9,11
untapped water sources. Management of wastewater from When safe discharge or reuse is not possible, brines are
industrial sources is particularly challenging as it can be highly commonly disposed of via deep well injection or treated using
saline, vary temporally in volume and composition, and contain evaporation ponds prior to the disposal of solids in
complex mixtures of contaminant species.3 The beneficial reuse landfills.9,11−13 These processes, however, have serious
of industrial wastewaters mitigates unsafe disposal by environmental limitations. For instance, some geologic
eliminating wastewater discharge, while simultaneously de- formations, such as those surrounding the Marcellus shale
creasing freshwater withdrawals.4,5 A major challenge for the fields in the United States, are unsuitable for wastewater
reuse of high-salinity wastewaters, however, is the high disposal via deep well injection because of the high potential to
energetic cost of brine desalination.6 In the case of inland contaminate groundwater.14 Furthermore, deep well injection
desalination, brine disposal costs often render the desalination has been linked to groundwater contamination in other areas
of brackish groundwater economically unfeasible.7 Low-cost as well as increased seismic activity.13 Volume minimization
brine management is critically needed to enable the utilization using evaporation ponds is similarly problematic as it poses a
of inland brackish groundwater and prevent the unsafe threat to groundwater and birdlife, requires large areas of land,
discharge of saline industrial wastewaters. and is effective in only warm, arid climates.15,16 The challenges
Because of the energy requirements for high-salinity associated with brine management often result in exorbitant
wastewater desalination, the preferred brine management disposal costs, which can even lead to otherwise promising
option, when it is both safe and effective to do so, is to brackish water resources remaining untapped as sources for
dispose or reuse the waste stream without extensive treatment. potable water.7 To avoid expensive brine disposal, an
For example, brine solutions (also called the retentate or
reject) produced from seawater reverse osmosis (SWRO) Received: May 26, 2018
processes can be safely discharged to the ocean without Revised: June 22, 2018
extensive treatment.8 In the oil and gas industry, produced Accepted: June 29, 2018
water undergoes relatively moderate treatment via dissolved air Published: June 29, 2018

© 2018 American Chemical Society 467 DOI: 10.1021/acs.estlett.8b00274


Environ. Sci. Technol. Lett. 2018, 5, 467−475
Environmental Science & Technology Letters Review

increasing level of attention is being given to brine RĜ (c P) + (1 − R )Ĝ (c R ) − Ĝ (c F)


concentration and zero liquid discharge (ZLD) practices, in SECmin =
R (1)
which waste is disposed of in solid form.17
A critical step in the ZLD process chain is desalination, where Ĝ is the specific Gibbs free energy as a function of
wherein water is recovered from a saline waste stream. composition at a fixed temperature and a fixed pressure and cP
Desalination can be achieved by membrane-based processes, and cR are the solute concentrations in the product and
such as reverse osmosis (RO) and electrodialysis, or phase- retentate streams, respectively. Assuming complete solute
change-based (designated here as “thermal”) processes, rejection, cP = 0 and cR = cF/(1 − R). Because of its salinity
including multi-effect distillation (MED), multi-stage flash dependence, SECmin is high for hypersaline brines (5.3 kWh
(MSF), and mechanical vapor compression (MVC).18 RO m−3 for 125000 mg L−1 at 50% recovery) as compared to those
utilizes hydraulic pressure, in excess of solution osmotic of more dilute solutions (1.1 kWh m−3 for 35000 mg L−1 at
pressure, to drive the transport of water across a semi- 50% recovery). In addition, the desalination energy efficiency,
permeable membrane while retaining most solutes.19 A η, is defined as the useful specific desalination work performed
combination of membrane and thermal processes are often by the system (i.e., SECmin) divided by its specific energy
used to achieve ZLD for saline wastewaters.20 First, RO consumption (SEC): η = SECmin/SEC.
concentrates wastewater to approximately 70000 mg L−1 total Membrane-based desalination processes such as RO and
dissolved solids (TDS), which has an osmotic pressure, π, of electrodialysis, which do not require a phase change to separate
∼59 bar. At higher salinities, thermal technologies are used to water from dissolved solutes, can achieve energy efficiencies of
concentrate brine streams to approximately 250000 mg L−1 (π >40%, particularly with multi-stage systems.24−26 In contrast,
≈ 290 bar), the typical inlet concentration for crystallizers in phase-change-based or thermal desalination processes, such as
ZLD.17 Thermal-based brine crystallizers then concentrate the MED, MSF, and MVC, are inherently less efficient, with typical
waste stream above its solubility limit (e.g., 357000 mg L−1 for η values of <20%.24,27 The energetic performance of thermal
NaCl) to extract solid salts for disposal.17 processes is strongly dependent on the recovery of the latent
Because of its superior energy efficiency, RO has displaced heat of vaporization, ΔvH. Given that ΔvH, >630 kWh m−3 for
thermal processes in recent decades for the major applications saline waters, is two orders of magnitude larger than the typical
of seawater and brackish water desalination for drinking water SECmin, imperfect heat recovery severely limits the energy
production.21 As such, membrane materials and processes have efficiency of all thermal desalination processes.27
To directly compare membrane-based and thermal-based
been optimized to treat feedwaters with salinities equal to or
desalination (Figure 1A), we can quantitatively analyze the
less than that of seawater (typically ∼35000 mg L−1; π ≈ 28
energetics of relevant processes. Typical SEC values and feed
bar). Consequently, the maximum operating pressure for RO is
concentrations are shown in Figure 1B for conventional
typically around 80 bar,22 suitable for overcoming the retentate (largely thermal) desalination technologies,17,27 in addition to
osmotic pressure of seawater treated to 50% recovery (∼70000 SEC calculated for HPRO. The SEC of HPRO is modeled
mg L−1; π ≈ 59 bar). Because of hydraulic-pressure limitations assuming a terminal hydraulic pressure (ΔPt) that is 5 bar
of RO, hypersaline brines, here defined as solutions with a above the osmotic pressure of the retentate (πR).26,28,29 More
TDS concentration of >70000 mg L −1, are primarily detailed insights can be gained by modeling different processes
desalinated via thermal processes.17 These processes, however, for the same inlet and outlet streams. In particular, a feed
are energy- and cost-intensive, which limits the implementa- solution of 70000 mg L−1 NaCl is concentrated to 250000 mg
tion of brine wastewater desalination prior to disposal. L−1 via HPRO-based, MVC-based, and hybrid HPRO−MVC
In this review, we critically discuss the application of high- desalination systems (Figure 1A). MVC is chosen as a
pressure RO (HPRO), defined here as RO operating above representative thermal process because of its highly effective
∼100 bar, to the treatment of hypersaline brines. We start by heat recovery. Its SEC is calculated assuming isentropic (i.e.,
considering the large volumes of industrial brines that could adiabatic and reversible) vapor compression with a terminal
require desalination and the energy requirements of HPRO temperature difference (ΔTt) of 5 °C above the boiling point
versus other processes. We largely consider two pressure limits elevation of the retentate (δR).24,30,31 In the two-stage HPRO
for HPRO: (i) 150 bar as roughly double the current operating and MVC models, water recovery in each stage is optimized to
limit and (ii) 300 bar to enable retentate concentrations of minimize SEC. In the hybrid HPRO−MVC model, the HPRO
250000 mg L−1, the inlet concentration for brine crystallizers. stage is limited to a maximum hydraulic pressure of 150 bar.
We then discuss the current state of HPRO and the process Further details of SEC calculations are given in the Supporting
design requirements for its effective application. We conclude Information.
with a discussion of the research needs that must be addressed Panels C and D of Figure 1 show the SEC and energy
to enable viable HPRO processes. efficiency of each process, which illustrates the drastic 3-fold


energy savings made possible using HPRO rather than a
ENERGY EFFICIENCY OF MEMBRANE thermal desalination process such as MVC. For example,
DESALINATION concentration of a 70000 mg L−1 hypersaline feed to 250000
mg L−1 would require 24 kWh m−3 with two-stage MVC (15%
Desalinating highly saline waste streams is inherently energy- energy efficiency) and only 7.3 kWh m−3 (47% energy
intensive. The minimum specific energy consumption efficiency) with two-stage HPRO. For context, the RO stage
(SECmin) of desalination (i.e., the energy required by a in a typical SWRO process can operate as low as 1.8 kWh m−3
thermodynamically reversible process per unit volume of (59% energy efficiency).25 Achieving a retentate concentration
product water) depends on the salinity of the feed stream, cF, of 250000 mg L−1 with HPRO would require approximately
and the water recovery ratio, R, the proportion of water 300 bar of hydraulic pressure, a challenging target from
recovered from the feed stream:23,24 membrane materials and module design perspectives. For this
468 DOI: 10.1021/acs.estlett.8b00274
Environ. Sci. Technol. Lett. 2018, 5, 467−475
Environmental Science & Technology Letters Review

Figure 1. (A) Brine desalination process schematics. Three different process configurations are shown for the concentration of a 70 g L−1 NaCl
solution to 250 g L−1. (B) Specific energy consumption (SEC) of various desalination processes. The typical feed concentration range and energy
consumption are shown for conventional RO and thermal brine concentrators.17,27 The expected SEC calculated for high-pressure reverse osmosis
(HPRO), at pressures of ≤150 and ≤300 bar, is also shown. (C) Specific energy consumption and (D) energy efficiency for HPRO and mechanical
vapor compression (MVC) brine desalination processes. The energy consumption requirements of five membrane-based and thermal-based
process configurations are analyzed for the desalination of 70 g L−1 NaCl at increasing retentate concentrations (i.e., water recovery). The energy
efficiency, η, is the ratio of the minimum energy of desalination for a given brine salinity and recovery, SECmin, to the SEC for each process. Shown
here are one- and two-stage processes for both HPRO and MVC. A hybrid HPRO−MVC process is also shown where an HPRO stage is employed
at a maximum applied pressure of ≤150 bar followed by an MVC stage. The water recovery ratio of each stage is optimized to reduce specific
energy consumption for each two-stage process. Further details about the calculations are provided in the Supporting Information.

reason, a hybrid process is modeled in which HPRO is limited regulations. In many circumstances, regulatory, financial, and
to a maximum pressure of 150 bar, after which the retentate is environmental restrictions motivate brine concentration and
further concentrated by MVC. At a retentate concentration of volume reduction.17 The coal-to-chemicals industry in China,
250000 mg L−1, the hybrid process SEC is 12 kWh m−3, for example, requires ZLD for all new facilities.32 Additionally,
representing a 2-fold reduction compared to that of two-stage the financial cost of brine disposal from inland desalination
MVC. Below a retentate osmotic pressure of 150 bar, the SEC facilities motivates brine volume minimization.7 In addition,
of the hybrid HPRO−MVC process is identical to the SEC of strict disposal regulations for landfill leachate in the United
one-stage HPRO, as the optimal water recovery ratio of its first States and effluent from the textile industry in India necessitate
stage (HPRO) is almost equal to its final recovery. As the brine volume reduction to minimize environmental impacts
retentate osmotic pressure increases beyond 150 bar, the water and health risks, despite the financial and energetic cost of
recovery of the second stage (MVC) begins to increase, brine desalination.33,34
leading to an increase in SEC. It is important to note that the Figure 2A shows the global desalination capacity that is
HPRO−MVC hybrid is a two-stage process and thus the brine currently installed or under construction for brine feedwaters
flow rate entering the less efficient MVC stage is significantly with a salinity of ≥50000 mg L−1 TDS.35 At present, these
lower than the initial flow rate (approximately 50% lower for brine desalination processes utilize conventional thermal
the conditions modeled). Consequently, the energetic separation processes (e.g., MVC, MED, or MSF) or conven-
performance of the hybrid HPRO−MVC process should be tional RO.35 Of the nearly 16000 desalination plants online or
assessed relative to two-stage HPRO and two-stage MVC, under construction in the GWI/IDA Desalting Inventory, only
rather than one-stage HPRO. 110 treat brine feedwaters.35 In practice, the high cost of brine


desalination limits hypersaline wastewater desalination before
disposal. Therefore, the total volume of hypersaline wastewater
POTENTIAL IMPACT OF HPRO present in each industry is surely much greater. In the United
A variety of industrial sources generate hypersaline brine States, the Environmental Protection Agency cites the high
wastewaters. Table 1 shows several brines of particular concern financial cost of brine desalination as the reason it does not
because of their large volume, concentration, or strict disposal require evaporative brine concentration of flue gas desulfuriza-
469 DOI: 10.1021/acs.estlett.8b00274
Environ. Sci. Technol. Lett. 2018, 5, 467−475
Environmental Science & Technology Letters Review

Table 1. Characteristics of Significant Brine Sources, Current Practices, and Disposal Regulations
representative flow typical TDS
brine wastewater rate (mg L−1) current disposal practice regulations regarding disposal
oil and gas produced water United States:11 13000−2100009 offshore: direct ocean discharge11 United States: regulations vary region-
8.7 million m3 day−1 onshore: deep well injection (where possible), ally based on permits9,11
reuse for hydraulic fracturing, evaporation Pennsylvania: discharge monthly
ponds9,11 average TDS of <500 mg L−173

brackish groundwater desali- United States:35,56 5000−5500056 surface or sewer discharge, United States: regulations vary region-
nation retentate 1.6 million m3 day−1 deep well injection (where possible)74 ally based on permits74

flue gas desulfurization United States:36 16000−5000036,75 settling ponds, chemical precipitation and United States: monthly average TDS of
(FGD) wastewater 1.2 million m3 day−1 surface discharge, zero liquid discharge36,76 <24 mg L−1 (effective starting No-
vember 2020)77,78

landfill leachate United States:79 0−5000079 zero liquid discharge via land application or United States: strict discharge regula-
230000 m3 day−1 recirculation to the landfill79 tions on 9 to 14 parameters not
including TDS33

coal-to-chemicals wastewater China:32 2000−1600017 zero liquid discharge32 China: zero liquid discharge required
320000 m3 day−1 for new plants17,32

textile industry wastewater estimates not available 1500−3000080 zero liquid discharge81 or chemical and India: TDS of <2100 mg L−1 for inland
biological treatment prior to surface discharge82 discharge34

Figure 2. (A) Global installed desalination capacity for brine feedwaters with total dissolved solids (TDS) concentrations of ≥50000 mg L−1.35 For
each industry, hollow bars represent the capacity desalinated by membrane-based processes and solid bars the brine capacity desalinated by thermal
processes. (B) Energy consumption of brine desalination. The left-most portion of the figure shows the daily energy consumption required to
desalinate the total quantity of brine wastewaters currently treated in each industry. The right-most portion of the figure shows a hypothetical
scenario in which the total volume of each saline wastewater presented in Table 1 must be concentrated to 250000 mg L−1 prior to disposal. In each
case, the energy consumption shown is that to concentrate the saline wastewaters to 70000 mg L−1 using a two-stage RO system followed by either
a two-stage HPRO or two-stage MVC process to concentrate the brine to 250000 mg L−1. Hollow bars represent the HPRO process scheme and
solid bars MVC. The corresponding percentage of the daily U.S. national electricity generation (NEG) is also shown.37 Additional calculation
details are provided in the Supporting Information.

tion (FGD) wastewater.36 If cost-effective technologies were approximately 1 million kWh day−1 (the equivalent of ∼34000
available, it is possible regulations would require brine U.S. homes)37 would be saved using HPRO as compared to
desalination prior to disposal. MVC.
At present, thermal technologies account for 54% of the Because of the lack of widespread brine desalination, a
global brine desalination capacity.35 The remaining 46% is hypothetical scenario (Figure 2B) is imagined to estimate the
treated using RO, but information about water recoveries and potential energy savings of HPRO compared to MVC for each
operating pressures is unavailable. To understand the energetic wastewater in Table 1. In this scenario, regulations require all
cost of treating these waste streams, we estimated the energy saline wastewaters to be concentrated to 250000 mg L−1 prior
(Figure 2B) needed to concentrate the total quantity of brine to disposal (i.e., approaching ZLD). The potential energy
wastewaters currently desalinated in each industry to 250000 savings are most notable for large-volume applications like
mg L−1 (initial volumes and concentrations are provided in produced water, where MVC would require 130 million kWh
Table S1). In these calculations, two-stage RO concentrates day−1, representing 1.2% of the daily U.S. NEG.37 In contrast,
wastewater up to 70000 mg L−1 followed by two-stage HPRO HPRO would utilize 85 million kWh day−1 less than MVC,
or two-stage MVC. The energy required to desalinate each albeit still requiring 0.4% of the daily U.S. NEG. 37
stream is shown alongside the proportion this represents of the Consequently, despite HPRO being 2−3-fold more efficient
average daily U.S. national electricity generation (NEG).37 In than MVC, its high thermodynamic minimum energy will
each scenario, MVC would require approximately 2−3 times mean considerable quantities of energy would be required for
more energy than HPRO. In the power industry, for example, HPRO. As such, work should be done to model the financial
470 DOI: 10.1021/acs.estlett.8b00274
Environ. Sci. Technol. Lett. 2018, 5, 467−475
Environmental Science & Technology Letters Review

costs and human health and environmental impacts of


hypersaline wastewater management38 to identify applications
■ HPRO PROCESS DESIGN CONSIDERATIONS
The design and operation of conventional RO processes are
best suited for HPRO. well-established for brackish water RO and SWRO. However,

■ CURRENT STATE OF MEMBRANE BRINE


DESALINATION
at high pressures and salinities, HPRO will likely require
unique design considerations. In this section, we discuss
several aspects of HPRO process design that must be
A few commercially available membranes rated above 80 bar addressed for the development of membrane-based hypersaline
brine desalination.
currently exist. Most notable is the Pall Disc Tube (DT)
Concentration polarization (CP) in the feed channel during
Module System, which has been used for landfill leachate
RO increases the osmotic pressure at the membrane surface,
treatment.39 DT modules are plate-and-frame configuration
πm, thus increasing the pressure needed to effect flux. Using
with large feed channels designed to reduce fouling from highly film theory and the van’t Hoff approximation, the CP modulus
contaminated leachate streams. As a result, they have a small can be defined as19
iJ y
= expjjjj w zzzz
active membrane area (9 m2 per element)40 compared to that
of high-pressure membranes with a spiral wound module πm − πp
kk{
design (27 m 2 per element). 41 Several DT module πb − π p (2)
configurations are available with a range of maximum pressure
during filtration from 70 to 150 bar.40 Average water output, where πb and πp are the osmotic pressures in the bulk feed and
however, is fairly low at 3 m3 day−1 per module for the permeate solutions, respectively, Jw is the water flux, and k is
treatment of landfill leachate or other wastewaters.40 the mass transfer coefficient. If we neglect the permeate
Membranes were also designed by Toray in the early 2000s osmotic pressure (πp = 0 assuming complete solute rejection)
for operation up to 100 bar to achieve 60% recovery SWRO.42 and changes in mass transfer coefficient, which will be affected
Nonetheless, at present, the maximum operating pressure for to some extent by salinity (Figure S1), the πm/πb ratio is
currently available Toray RO membranes is 83 bar.43 constant for a given water flux, meaning that the relative
Additionally, Dow recently started manufacturing a specialty increase in osmotic pressure should be similar for different feed
line of RO elements rated at operating pressures up to 120 solutions. Conversely, the absolute increase in osmotic
bar.41,44 The largest of these spiral wound elements (8 in. pressure from CP is correspondingly greater for high-salinity
diameter) is rated to generate permeate flow rates of 24.2 m3 feeds. For example, a modest CP modulus of 1.15 increases the
day−1, albeit when treating 32000 mg L−1 NaCl to 8% recovery feed osmotic pressure of flowback water (modeled as 157000
at 55 bar (standard SWRO test conditions).41 The achievable mg L−1 TDS)50 by 26 bar, approximately the osmotic pressure
permeate flow rate during high-pressure operation is unclear. of seawater.
To circumvent the need for high hydraulic pressures, We performed module-scale modeling to assess the impact
alternative “osmotically assisted” RO process configurations of CP for one- and two-stage HPRO processes (Figure 3A).
that theoretically could use conventional (∼70 bar) pressures When comparing 35000, 70000, and 125000 mg L−1 feeds
have recently been proposed.45,46 These hypothetical processes (each at 50% recovery), we find that water fluxes are
would use a saline solution at the membrane permeate side to surprisingly similar. This result is partially a function of
reduce the osmotic pressure difference across the membrane conventional direct-pass RO operation where the hydraulic
and thus the required applied hydraulic pressure.45,46 Several pressure driving force is roughly the same throughout the
stages would be used to sequentially decrease the feed and module: the increased driving force at the head of the module
permeate osmotic pressures, until finally conventional RO can for hypersaline feeds roughly balances the impact of increased
be used to produce pure water. While these osmotically CP. To improve process efficiency, interstage design (i.e.,
assisted RO processes could theoretically avoid high-pressure placing higher-rejection, low-flux membrane elements at the
head of the module) may be useful for balancing water flux
operation, their performance is inherently limited by internal
along the module length (Figure S2).51 While this analysis
concentration polarization (ICP) in the membrane support
considers only a few cases, it appears that SWRO-like water
layer. As previous studies on forward osmosis have shown, ICP
fluxes (10−15 L m−2 h−1) should theoretically be attainable for
decreases the solute concentration at the permeate-side HPRO, despite the increased CP stemming from the high feed
membrane−solution interface, drastically reducing the trans- osmotic pressures.
membrane driving force for water permeation and thus limiting Although CP is expected to impact water flux relatively little
water flux.47,48 in HPRO, it may critically influence inorganic scaling. In RO,
Designing membranes for osmotically assisted RO to sparingly soluble salts near their solubility limit can precipitate
mitigate ICP will be very challenging. An intrinsic trade-off because of CP or increased retentate concentrations as water
exists between ICP reduction and mechanical integrity: the recovery increases.52 Scaling could be a performance-limiting
membrane properties that decrease ICP (small support layer phenomenon in HPRO processes, particularly at high water
thickness and high porosity) will decrease the pressure recoveries (e.g., ZLD). Common scale-forming species,
tolerance of the membrane. The use of pressure-retarded including Ca2+, Mg2+, CO32−, SO42−, Ba2+, Sr2+, and silica,53,54
osmosis (PRO) membranes has been proposed to balance are present in many hypersaline brine wastewaters. FGD
these contradictory needs;45 however, recent work has shown effluent, for example, has average concentrations of gypsum
dramatic deformation of PRO membranes to take place at an precursors Ca2+ and SO42− of 2000 and 13000 mg L−1,
applied pressure of 55 bar, the maximum pressure achieved respectively.36 If FGD effluent (33000 mg L−1 TDS)36 was
prior to failure.49 In addition to ICP, increased capital costs concentrated to 250000 mg L−1 TDS (87% recovery), Ca2+
from staging and high frictional pressure drops in the permeate and SO42− concentrations would be 15000 and 100000 mg
channels49 would also hinder the proposed processes. L−1, respectively, far above the 2400 mg L−1 solubility limit of
471 DOI: 10.1021/acs.estlett.8b00274
Environ. Sci. Technol. Lett. 2018, 5, 467−475
Environmental Science & Technology Letters Review

Figure 3. (A) Average water flux calculated from module-scale membrane process modeling. The average water flux for one- and two-stage HPRO
processes at 50% water recovery is shown for different terminal hydraulic pressures, ΔPt (i.e., applied pressure in excess of the retentate osmotic
pressure). Osmotic pressures of the exiting retentate solution are also shown. Mass transfer coefficients are calculated using Sherwood−Reynolds−
Schmidt correlations and a water permeability coefficient, A, of 1 L m−2 h−1 bar−1 is assumed. (B) Specific membrane area for common module
designs. The specific membrane area is shown as a function of characteristic length along with the permeate production rate for a typical 8 in.
diameter, 40 in. long module operating at a water flux of 10 L m−2 h−1. The characteristic length shown here is the feed channel height for typical
spiral wound68 and plate-and-frame modules.69 For hollow fiber modules, the characteristic length is the average distance between fibers for typical
module packing densities.19 In the schematics, dark blue denotes the feed stream, light blue the permeate, black the membrane active layer, and
gray the membrane support layer, and the characteristic length, a, is colored white. (C) Compression pressure for nonporous hollow fibers. The
compression pressure is calculated using the von Mises criterion as a function of the fiber radii ratio.66 Shown here are dense, nonporous materials,
which will have a compressive yield strength greater than that of porous hollow fiber membranes. The materials shown are polytetrafluoroethylene
(PTFE), cellulose acetate (CA), polyvinyl chloride (PVC), polysulfone (PSf), polyimide (PI), A36 Steel, and Kevlar 49. The compressive yield
strength, σ, is listed for each material.70−72

gypsum.55 In addition, antiscalants that are commonly used to waters up to 70000 mg L−1 due to capital cost concerns. At
prevent scaling in conventional RO53 have been ineffective at higher concentrations (i.e., HPRO), energy costs would likely
high ionic strengths.56 Pretreatment processes are commonly incentivize two-stage processes, which would substantially
employed to remove foulants in conventional RO desalination reduce energy requirements compared to those of a one-stage
processes.22 Similarly, pretreatment to remove scale-forming process, as shown in Figure 1C. Additional configurations are
species may be necessary to achieve high water recoveries in also possible such as batch and semibatch RO (also called
HPRO. closed-circuit RO), which use retentate recycle streams.26
An important consideration for HPRO is the large volumes Energetic modeling has shown these processes could achieve
of hypersaline brines that exist globally. To meet this need, similar energy consumption as two- or three-stage RO, without
HPRO technologies should have a large specific membrane physical staging.26


area (i.e., membrane surface area per unit volume of membrane
module) to desalinate large-volume brines in compact facilities. RESEARCH NEEDS FOR HPRO
Figure 3B shows the specific membrane surface areas for
several common module design configurations. While hollow Conventional RO is a mature technology with high-performing
fibers would provide the largest specific areas, their employ- materials, well-understood transport phenomena, and robust
ment in RO is affected by high bore-side pressure drops and process designs.25 However, much is unknown about the
poorly defined shell-side fluid flow.19 In contrast, spiral wound behavior of membrane materials and processes under HPRO
modules are commonly used for conventional RO, and because conditions. In this section, we highlight the primary research
SWRO-like water fluxes are expected for HPRO (Figure 3A), needs for the development and large-scale implementation of
they are potentially an optimal HPRO module design. The HPRO.
design of modules and pressure vessels suitable for HPRO One of the greatest unknown aspects of HPRO is the
should be relatively simple using high-strength materials or fundamental effect of membrane compaction (i.e., physical
more robust designs. Costs may increase for auxiliary system deformation) on membrane performance at high pressure. At
components and piping that must withstand high pressure and present, only a few studies have assessed RO above 100 bar,
resist corrosion when exposed to hypersaline brines. The most with maximum tested pressures of 200 bar.57,58 These studies
challenging module design consideration is likely gluing observed decreased water permeability at high pressure and
membrane sheets together to form an envelope that can attributed it to compaction, although compaction was not
withstand high pressure, although this can likely be achieved directly observed. Indeed, even at conventional RO pressures,
using high-strength adhesives. Permeate spacers should also be much remains unknown about the nature of compaction.
designed for HPRO to minimize membrane deformation into Laboratory-scale studies commonly attribute declining water
spacer voids without large increases in frictional pressure permeability to compaction; however, recent work indicates
losses. fouling may be the primary cause.59 Compaction is best
From a process design standpoint, high salinities in HPRO studied in ultrafiltration (UF) processes, where it has a strong
will incentivize unconventional configurations to decrease dependence on membrane composition and morphology.60−63
energy consumption. Conventional RO process designs (e.g., Decreased water permeability has been correlated to changes
one-stage RO) are likely the best options to concentrate feed in overall thickness,63 and some results speculate it is primarily
472 DOI: 10.1021/acs.estlett.8b00274
Environ. Sci. Technol. Lett. 2018, 5, 467−475
Environmental Science & Technology Letters Review

a result of compaction in the uppermost skin layer of UF constant for a given membrane.19 However, recent work has
membranes.62 The compaction behavior of UF membranes found B to increase substantially with salt concentration,
may contribute to developing an understanding of RO possibly because of the decreased impacts of electrostatic
compaction as the support layer of state-of-the-art thin-film repulsion to reject ions at higher salinities.49 To understand
composite (TFC) membranes is typically a polysulfone (PSf) this behavior, the mechanisms and assumptions of the
UF membrane. However, the polyamide selective layer solution−diffusion model must be studied under HPRO
dominates resistance in TFC membranes, which complicates conditions.
direct translation from the study of UF membranes. In TFC
membranes, compaction is typically considered to occur in the
support layer;64 however, a recent study speculated that
■ OUTLOOK FOR HPRO
The development of HPRO has the potential to enable energy-
compaction can occur in the dense selective layer, as well.65 A efficient and cost-effective hypersaline brine desalination,
much greater understanding of the location and impact of although, because of the high thermodynamic minimum
compaction is critically needed for the development of energy of desalination, the quantity of energy required by
deformation-resistant HPRO membranes. HPRO remains relatively large. Therefore, HPRO may be best-
Once the mechanisms of compaction are understood, high- suited for applications where strict regulations or operational
strength membranes must be fabricated with high water circumstances preclude direct reuse or safe disposal of saline
permeability and salt rejection under HPRO conditions. At brines. Nevertheless, the discharge of saline wastewater to the
high pressures, membranes may follow two possible modes of environment is undesirable. If the development of HPRO
failure: severe compaction, resulting in no permeability, or enables energy-efficient brine desalination, future regulations
rupture, resulting in no salt rejection. In the case of severe may further restrict brine discharge practices, motivating
compaction, polymer networks may deform to the extent they greater need for ZLD.
form a dense, impermeable film. Above a certain pressure limit, Decades of work improving conventional RO will provide a
however, membranes are likely to rupture, causing a foundational knowledge from which HPRO can be developed.
catastrophic loss of solute rejection. In spiral wound It is possible that many aspects of conventional RO membrane
membranes, rupture is likely to occur within the voids of the and process design will directly translate to HPRO. Even so,
permeate-side spacer. Alternatively, in hollow fibers with a significant work will be needed where knowledge gaps exist,
shell-side feed, membranes are likely to collapse under notably the fundamental nature of compaction. Similarly,
pressure. inorganic scaling, a performance-limiting phenomenon in
To gain an understanding of membrane mechanical strength, conventional RO, may limit the achievable recoveries of
Figure 3C estimates the compression pressure for nonporous HPRO and necessitate the development of effective and
hollow fibers fabricated from different polymers, with steel and efficient pretreatment to remove scale-forming species.
Kevlar shown for reference. This analysis focuses on hollow Current brine management practices are expensive and
fiber membranes because their cylindrical shape is more well- environmentally unsustainable. At present, saline waste streams
defined than flat sheet membranes supported by permeate are discharged to the environment, evaporated in large ponds,
spacers, as in spiral wound modules. The von Mises criterion or injected into the ground. Where ZLD is required, inefficient
estimates the first point of fiber compression as a function of thermal separation processes must be used. These costly and
applied pressure, material compressive yield strength, and fiber environmentally unsound disposal practices prevent the use of
radii ratio (details in the Supporting Information).66,67 The inland brackish groundwater, often abundant in water scarce
results in Figure 3C, calculated for nonporous polymer regions, and inhibit the beneficial reuse of saline industrial
materials, suggest high pressures may be achievable; however, wastewater. To help increase the utilization of these waters,
the material strength will decrease considerably for porous HPRO offers great promise as an energy-efficient and cost-
materials. A critical research need is to identify porous effective brine desalination technology.
membrane morphologies that exhibit high compressive yield
strength without compromising membrane performance. It can
be seen that polysulfone (PSf), the typical TFC RO support

*
ASSOCIATED CONTENT
S Supporting Information
layer material, has a relatively large pressure tolerance in The Supporting Information is available free of charge on the
comparison to those of the other polymers shown. As such, it ACS Publications website at DOI: 10.1021/acs.es-
may be necessary to develop novel high-strength support layer tlett.8b00274.
materials to provide the pressure resistance needed for HPRO. Derivation of energy consumption equations, method-
Figure 3C also shows a plateau in compression pressure may ology of module-scale water flux analysis, methodology
be reached at a certain fiber radii ratio, indicating that an
for specific area calculations of various module designs,
increased support layer thickness may not significantly increase
calculation of nonporous hollow fiber compression
pressure resistance and high-strength materials may be needed.
pressure, industrial wastewater data (Table S1), depend-
In addition to the unknown behavior of compaction,
ence of the CP modulus on Re and Sc (Figure S1), water
transport mechanisms in HPRO are poorly understood.
Water and solute transport in RO is traditionally characterized flux as a function of module position (Figure S2), and
by the solution−diffusion model.19 While this model is widely pressure drop for various feed concentrations (Figure
accepted and adequate for conventional RO, it has not been S3) (PDF)
tested at salinities and pressures relevant to HPRO. In the
solution−diffusion model, salt flux, Js, is typically defined as Js =
BΔC, where ΔC is the difference in salt concentration between
■ AUTHOR INFORMATION
Corresponding Author
the feed- and permeate-side interfaces of the selective layer and *E-mail: menachem.elimelech@yale.edu. Phone: (203) 432-
B is the salt permeability coefficient, typically assumed to be 2789.
473 DOI: 10.1021/acs.estlett.8b00274
Environ. Sci. Technol. Lett. 2018, 5, 467−475
Environmental Science & Technology Letters Review

ORCID (18) Bahar, R.; Hawlader, M. N. A.; Woei, L. S. Performance


Menachem Elimelech: 0000-0003-4186-1563 evaluation of a mechanical vapor compression desalination system.
Desalination 2004, 166, 123−127.
Notes (19) Baker, R. W. Membrane Technology and Applications.; McGraw-
The authors declare no competing financial interest. Hill: New York, 2000.

■ ACKNOWLEDGMENTS
The authors acknowledge the support received from the
(20) Stanford, B. D.; Leising, J. F.; Bond, R. G.; Snyder, S. A. Inland
Desalination: Current Practices, Environmental Implications, and
Case Studies in Las Vegas, NV. In Sustainability Science and
Engineering; Escobar, I. C., Schäfer, A. I., Eds.; Elsevier: Amsterdam,
National Science Foundation under Grant CBET-1701658 and 2010; Vol. 2, pp 327−350.
Graduate Research Fellowship DGE-1122492 awarded to (21) Desalination Markets 2016; Global Water Intelligence: Oxford,
J.R.W. U.K., 2016.

■ REFERENCES
(1) Water for a Sustainable World. United Nations Educational,
(22) Fritzmann, C.; Löwenberg, J.; Wintgens, T.; Melin, T. State-of-
the-art of reverse osmosis desalination. Desalination 2007, 216, 1−76.
(23) Mistry, K.; Lienhard, J. Generalized Least Energy of Separation
for Desalination and Other Chemical Separation Processes. Entropy
Scientific and Cultural Organization: New York, 2015.
2013, 15, 2046.
(2) A Post-2015 Global Goal for Water: Synthesis of key findings
(24) Mistry, K. H.; McGovern, R. K.; Thiel, G. P.; Summers, E. K.;
and recommendations from UN-Water. United Nations Water: New
York, 2014. Zubair, S. M.; Lienhard, J. H. Entropy Generation Analysis of
(3) Lefebvre, O.; Moletta, R. Treatment of organic pollution in Desalination Technologies. Entropy 2011, 13, 1829.
industrial saline wastewater: A literature review. Water Res. 2006, 40, (25) Elimelech, M.; Phillip, W. A. The Future of Seawater
3671−3682. Desalination: Energy, Technology, and the Environment. Science
(4) Grant, S. B.; Saphores, J.-D.; Feldman, D. L.; Hamilton, A. J.; 2011, 333, 712−717.
Fletcher, T. D.; Cook, P. L. M.; Stewardson, M.; Sanders, B. F.; Levin, (26) Werber, J. R.; Deshmukh, A.; Elimelech, M. Can batch or semi-
L. A.; Ambrose, R. F.; Deletic, A.; Brown, R.; Jiang, S. C.; Rosso, D.; batch processes save energy in reverse-osmosis desalination?
Cooper, W. J.; Marusic, I. Taking the “Waste” Out of “Wastewater” Desalination 2017, 402, 109−122.
for Human Water Security and Ecosystem Sustainability. Science (27) Deshmukh, A.; Boo, C.; Karanikola, V.; Lin, S.; Straub, A. P.;
2012, 337, 681−686. Tong, T.; Warsinger, D. M.; Elimelech, M. Membrane distillation at
(5) Adewumi, J. R.; Ilemobade, A. A.; Van Zyl, J. E. Treated the water-energy nexus: limits, opportunities, and challenges. Energy
wastewater reuse in South Africa: Overview, potential and challenges. Environ. Sci. 2018, 11, 1177−1196.
Resour., Conserv. and Recycl. 2010, 55, 221−231. (28) Sharqawy, M. H.; Lienhard, J. H.; Zubair, S. M.
(6) Zander, A. K.; Elimelech, M.; Furukawa, D. H.; Gleick, P.; Herd, Thermophysical properties of seawater: a review of existing
K. R.; Jones, K. L.; Rolchigo, P.; Sethi, S.; Tonner, J.; Vaux, H. J.; correlations and data. Desalin. Water Treat. 2010, 16, 354−380.
Weis, J. S.; Wood, W. W. Desalination: A National Perspective; (29) Nayar, K. G.; Sharqawy, M. H.; Banchik, L. D.; Lienhard V, J.
National Research Council: Washington, DC, 2008; p 312. H. Thermophysical properties of seawater: A review and new
(7) Brady, P. V.; Kottenstette, R. J.; Mayer, T. M.; Hightower, M. M. correlations that include pressure dependence. Desalination 2016,
Inland Desalination: Challenges and Research Needs. Journal of 390, 1−24.
Contemporary Water Research & Education 2005, 132, 46−51. (30) Thiel, G. P.; Tow, E. W.; Banchik, L. D.; Chung, H. W.;
(8) Lattemann, S.; Höpner, T. Environmental impact and impact Lienhard, J. H. Energy consumption in desalinating produced water
assessment of seawater desalination. Desalination 2008, 220, 1−15. from shale oil and gas extraction. Desalination 2015, 366, 94−112.
(9) Technical Development Document for the Effluent Limitations (31) Stoughton, R. W.; Lietzke, M. H. Thermodynamic properties of
Guidelines and Standards for the Oil and Gas Extraction Point Source sea salt solutions. J. Chem. Eng. Data 1967, 12, 101−104.
Category. U.S. Environmental Protection Agency: Washington, DC, (32) Xiong, R.; Wei, C. Current status and technology trends of zero
2016. liquid discharge at coal chemical industry in China. Journal of Water
(10) Bakke, T.; Klungsøyr, J.; Sanni, S. Environmental impacts of Process Engineering 2017, 19, 346−351.
produced water and drilling waste discharges from the Norwegian (33) Effluent Limitations Guidelines, Pretreatment Standards, and
offshore petroleum industry. Mar. Environ. Res. 2013, 92, 154−169. New Source Performance Standards for the Landfills Point Source
(11) Clark, C. E.; Veil, J. A. Produced water volumes and management Category; Final Rule. U.S. Environmental Protection Agency:
practices in the United States; Argonne National Laboratory: Argonne, Washington, DC, 2000.
IL, 2009. (34) Environment (Protection) Fifth Amendment Rules, 2016. In
(12) Shaffer, D. L.; Arias Chavez, L. H.; Ben-Sasson, M.; Romero- India Ministry of Environment. The Gazette of India: New Delhi,
Vargas Castrillón, S.; Yip, N. Y.; Elimelech, M. Desalination and 2016; Vol. G.S.R. 978(E).
Reuse of High-Salinity Shale Gas Produced Water: Drivers, (35) GWI/IDA Desalting Inventory. http://www.desaldata.com
Technologies, and Future Directions. Environ. Sci. Technol. 2013, (accessed February 26, 2018).
47, 9569−9583. (36) Technical Development Document for the Effluent Limitations
(13) Vidic, R. D.; Brantley, S. L.; Vandenbossche, J. M.; Yoxtheimer, Guidelines and Standards for the Steam Electric Power Generating
D.; Abad, J. D. Impact of Shale Gas Development on Regional Water Point Source Category. U.S. Environmental Protection Agency:
Quality. Science 2013, 340, 1235009. Washington, DC, 2015.
(14) Lutz, B. D.; Lewis, A. N.; Doyle, M. W. Generation, transport, (37) Electric Power Annual 2016. U.S. Energy Information
and disposal of wastewater associated with Marcellus Shale gas Administration: Washington, DC, 2017.
development. Water Resour. Res. 2013, 49, 647−656. (38) Bartholomew, T. V.; Mauter, M. S. Multiobjective Optimization
(15) Ahmed, M.; Shayya, W. H.; Hoey, D.; Mahendran, A.; Morris, Model for Minimizing Cost and Environmental Impact in Shale Gas
R.; Al-Handaly, J. Use of evaporation ponds for brine disposal in Water and Wastewater Management. ACS Sustainable Chem. Eng.
desalination plants. Desalination 2000, 130, 155−168. 2016, 4, 3728−3735.
(16) Ramirez, P. Bird Mortality in Oil Field Wastewater Disposal (39) Renou, S.; Givaudan, J. G.; Poulain, S.; Dirassouyan, F.;
Facilities. Environ. Manage. 2010, 46, 820−826. Moulin, P. Landfill leachate treatment: Review and opportunity. J.
(17) Tong, T.; Elimelech, M. The Global Rise of Zero Liquid Hazard. Mater. 2008, 150, 468−493.
Discharge for Wastewater Management: Drivers, Technologies, and (40) Disc Tube Module System Datasheet. Pall Corp.: Port
Future Directions. Environ. Sci. Technol. 2016, 50, 6846−6855. Washington, NY, 2010.

474 DOI: 10.1021/acs.estlett.8b00274


Environ. Sci. Technol. Lett. 2018, 5, 467−475
Environmental Science & Technology Letters Review

(41) DOW XUS180808 Reverse Osmosis Element Product Data (63) Aghajani, M.; Maruf, S. H.; Wang, M.; Yoshimura, J.; Pichorim,
Sheet. The Dow Chemical Co.: Midland, MI, 2016. G.; Greenberg, A.; Ding, Y. Relationship between permeation and
(42) Taniguchi, M.; Kurihara, M.; Kimura, S. Behavior of a reverse deformation for porous membranes. J. Membr. Sci. 2017, 526, 293−
osmosis plant adopting a brine conversion two-stage process and its 300.
computer simulation. J. Membr. Sci. 2001, 183, 249−257. (64) Jonsson, G. Methods for determining the selectivity of reverse
(43) Standard SWRO: TM800M. Toray Industries, Inc.: Tokyo, osmosis membranes. Desalination 1977, 24, 19−37.
2014. (65) Pendergast, M. T. M.; Nygaard, J. M.; Ghosh, A. K.; Hoek, E.
(44) DOW Specialty Membrane XUS180804 and XUS180802 M. V. Using nanocomposite materials technology to understand and
Reverse Osmosis Elements Product Data Sheet. The Dow Chemical control reverse osmosis membrane compaction. Desalination 2010,
Co.: Midland, MI, 2017. 261, 255−263.
(45) Bartholomew, T. V.; Mey, L.; Arena, J. T.; Siefert, N. S.; (66) Ekiner, O. M.; Vassilatos, G. Polyaramide hollow fibers for
Mauter, M. S. Osmotically assisted reverse osmosis for high salinity hydrogen/methane separation  spinning and properties. J. Membr.
brine treatment. Desalination 2017, 421, 3−11. Sci. 1990, 53, 259−273.
(46) Chen, X.; Yip, N. Y. Unlocking High-Salinity Desalination with (67) Koh, D.-Y.; McCool, B. A.; Deckman, H. W.; Lively, R. P.
Cascading Osmotically Mediated Reverse Osmosis: Energy and Reverse osmosis molecular differentiation of organic liquids using
Operating Pressure Analysis. Environ. Sci. Technol. 2018, 52, 2242− carbon molecular sieve membranes. Science 2016, 353, 804−807.
2250. (68) Schock, G.; Miquel, A. Mass transfer and pressure loss in spiral
(47) McCutcheon, J. R.; Elimelech, M. Influence of concentrative wound modules. Desalination 1987, 64, 339−352.
and dilutive internal concentration polarization on flux behavior in (69) Balster, J.; Pünt, I.; Stamatialis, D. F.; Wessling, M. Multi-layer
forward osmosis. J. Membr. Sci. 2006, 284, 237−247. spacer geometries with improved mass transport. J. Membr. Sci. 2006,
(48) Deshmukh, A.; Yip, N. Y.; Lin, S.; Elimelech, M. Desalination 282, 351−361.
by forward osmosis: Identifying performance limiting parameters (70) Inoue, N. Polymers. In Hydrostatic Extrusion: Theory and
through module-scale modeling. J. Membr. Sci. 2015, 491, 159−167. Applications; Inoue, N., Nishihara, M., Eds.; Springer: Dordrecht, The
(49) Straub, A. P.; Osuji, C. O.; Cath, T. Y.; Elimelech, M. Netherlands, 1985; pp 333−362.
Selectivity and Mass Transfer Limitations in Pressure-Retarded (71) Raddon, B. J. Biaxial restraint of axially loaded steel cores. M.S.
Osmosis at High Concentrations and Increased Operating Pressures. Thesis, The University of Utah, Salt Lake City, UT, 2010.
Environ. Sci. Technol. 2015, 49, 12551−12559. (72) Deteresa, S. J.; Allen, S. R.; Farris, R. J.; Porter, R. S.
(50) Haluszczak, L. O.; Rose, A. W.; Kump, L. R. Geochemical Compressive and torsional behaviour of Kevlar 49 fibre. J. Mater. Sci.
evaluation of flowback brine from Marcellus gas wells in Pennsylvania, 1984, 19, 57−72.
USA. Appl. Geochem. 2013, 28, 55−61. (73) Renner, R. Pennsylvania to regulate salt discharges. Environ. Sci.
(51) Peñate, B.; García-Rodríguez, L. Reverse osmosis hybrid Technol. 2009, 43, 6120−6120.
membrane inter-stage design: A comparative performance assessment. (74) Mickley, M. C. Membrane Concentrate Disposal: Practices and
Desalination 2011, 281, 354−363. Regulation. U.S. Department of the Interior Bureau of Reclamation:
(52) Shirazi, S.; Lin, C.-J.; Chen, D. Inorganic fouling of pressure- Washington, DC, 2006.
driven membrane processes  A critical review. Desalination 2010, (75) Huang, Y. H.; Peddi, P. K.; Tang, C.; Zeng, H.; Teng, X.
250, 236−248. Hybrid zero-valent iron process for removing heavy metals and nitrate
(53) Antony, A.; Low, J. H.; Gray, S.; Childress, A. E.; Le-Clech, P.; from flue-gas-desulfurization wastewater. Sep. Purif. Technol. 2013,
Leslie, G. Scale formation and control in high pressure membrane 118, 690−698.
water treatment systems: A review. J. Membr. Sci. 2011, 383, 1−16. (76) Gingerich, D. B.; Grol, E.; Mauter, M. S. Fundamental
(54) Fakhru’l-Razi, A.; Pendashteh, A.; Abdullah, L. C.; Biak, D. R. challenges and engineering opportunities in flue gas desulfurization
A.; Madaeni, S. S.; Abidin, Z. Z. Review of technologies for oil and gas wastewater treatment at coal fired power plants. Environ. Sci.: Water
produced water treatment. J. Hazard. Mater. 2009, 170, 530−551. Res. Technol. 2018, 4, 909−925.
(55) Deng, M.; Liu, Q.; Xu, Z. Impact of gypsum supersaturated (77) Effluent Limitations Guidelines and Standards for the Steam
water on the uptake of copper and xanthate on sphalerite. Miner. Eng. Electric Power Generating Point Source Category; Final Rule. U.S.
2013, 49, 165−171. Environmental Protection Agency: Washington, DC, 2015.
(56) Greenlee, L. F.; Lawler, D. F.; Freeman, B. D.; Marrot, B.; (78) Postponement of Certain Compliance Dates for the Effluent
Moulin, P. Reverse osmosis desalination: Water sources, technology, Limitations Guidelines and Standards for the Steam Electric Power
and today’s challenges. Water Res. 2009, 43, 2317−2348. Generating Point Source Category. U.S. Environmental Protection
(57) Rautenbach, R.; Linn, T. High-pressure reverse osmosis and Agency: Washington, DC, 2017.
nanofiltration, a “zero discharge” process combination for the (79) Browner, C. M.; Fox, J. C.; Grubbs, G. H.; Frace, S. E.; Forsht,
treatment of waste water with severe fouling/scaling potential. E. H.; Ebner, M. C. Development Document for Final Effluent
Desalination 1996, 105, 63−70. Limitations Guidelines and Standards for the Landfills Point Source
(58) Rautenbach, R.; Linn, T.; Eilers, L. Treatment of severely Category. U.S. Environmental Protection Agency: Washington, DC,
2000.
contaminated waste water by a combination of RO, high-pressure RO
(80) Dasgupta, J.; Sikder, J.; Chakraborty, S.; Curcio, S.; Drioli, E.
and NF  potential and limits of the process. J. Membr. Sci. 2000,
Remediation of textile effluents by membrane based treatment
174, 231−241.
techniques: A state of the art review. J. Environ. Manage. 2015, 147,
(59) Van Wagner, E. M.; Sagle, A. C.; Sharma, M. M.; Freeman, B.
55−72.
D. Effect of crossflow testing conditions, including feed pH and
(81) Vishnu, G.; Palanisamy, S.; Joseph, K. Assessment of fieldscale
continuous feed filtration, on commercial reverse osmosis membrane
zero liquid discharge treatment systems for recovery of water and salt
performance. J. Membr. Sci. 2009, 345, 97−109.
from textile effluents. J. Cleaner Prod. 2008, 16, 1081−1089.
(60) Brinkert, L.; Abidine, N.; Aptel, P. On the relation between
(82) Verma, A. K.; Dash, R. R.; Bhunia, P. A review on chemical
compaction and mechanical properties for ultrafiltration hollow fibers.
coagulation/flocculation technologies for removal of colour from
J. Membr. Sci. 1993, 77, 123−131.
textile wastewaters. J. Environ. Manage. 2012, 93, 154−168.
(61) Persson, K. M.; Gekas, V.; Trägårdh, G. Study of membrane
compaction and its influence on ultrafiltration water permeability. J.
Membr. Sci. 1995, 100, 155−162.
(62) Stade, S.; Kallioinen, M.; Mikkola, A.; Tuuva, T.; Mänttäri, M.
Reversible and irreversible compaction of ultrafiltration membranes.
Sep. Purif. Technol. 2013, 118, 127−134.

475 DOI: 10.1021/acs.estlett.8b00274


Environ. Sci. Technol. Lett. 2018, 5, 467−475

You might also like